Paul Kirk and Matthias Lesch - arXiv · Patodi–Singer ρ–invariant and relate it to Wall’s...

66
arXiv:math/0012123v2 [math.DG] 11 Mar 2002 The η –invariant, Maslov index, and spectral flow for Dirac–type operators on manifolds with boundary Paul Kirk and Matthias Lesch Abstract. Several proofs have been published of the mod Z gluing formula for the η–invariant of a Dirac operator. However, so far the integer contribution to the gluing formula for the η–invariant is left obscure in the literature. In this article we present a gluing formula for the η–invariant which expresses the integer contribution as a triple index involving the boundary conditions and the Calder´ on projectors of the two parts of the decomposition. The main ingredients of our presentation are the Scott– Wojciechowski theorem for the determinant of a Dirac operator on a manifold with boundary and the approach of Br¨ uning–Lesch to the mod Z gluing formula. Our presentation includes careful constructions of the Maslov index and triple index in a symplectic Hilbert space. As a byproduct we give intuitively appealing proofs of two theorems of Nicolaescu on the spectral flow of Dirac operators. As an application of our methods, we carry out a detailed analysis of the ηinvariant of the odd signature operator coupled to a flat connection using adiabatic methods. This is used to extend the definition of the Atiyah–Patodi–Singer ρ–invariant to manifolds with boundary. We derive a “non–additivity” formula for the Atiyah– Patodi–Singer ρ–invariant and relate it to Wall’s non-additivity formula for the sig- nature of even–dimensional manifolds. Contents 1. Introduction 2 2. Dirac operators on manifolds with boundary and the self–adjoint Fredholm Grassmannian 4 3. The η–invariant and spectral flow 11 4. The Scott–Wojciechowski theorem 13 5. Splittings of manifolds and the η–invariant I 17 6. Maslov index and winding number 25 7. Splittings of manifolds and the η–invariant II 35 8. Adiabatic stretching and applications to the Atiyah-Patodi-Singer ρ-invariant 39 References 64 The first named author gratefully acknowledges the support of the National Science Foundation under grant no. DMS-9971020. The second named author was supported by a Heisenberg fellowship of Deutsche Forschungsgemeinschaft and by the National Science Foundation under grant no. DMS- 0072551 . 1

Transcript of Paul Kirk and Matthias Lesch - arXiv · Patodi–Singer ρ–invariant and relate it to Wall’s...

arX

iv:m

ath/

0012

123v

2 [

mat

h.D

G]

11

Mar

200

2

The η–invariant, Maslov index, and spectral flow for

Dirac–type operators on manifolds with boundary

Paul Kirk and Matthias Lesch

Abstract. Several proofs have been published of the modZ gluing formula for theη–invariant of a Dirac operator. However, so far the integer contribution to the gluingformula for the η–invariant is left obscure in the literature. In this article we presenta gluing formula for the η–invariant which expresses the integer contribution as atriple index involving the boundary conditions and the Calderon projectors of the twoparts of the decomposition. The main ingredients of our presentation are the Scott–Wojciechowski theorem for the determinant of a Dirac operator on a manifold withboundary and the approach of Bruning–Lesch to the modZ gluing formula.

Our presentation includes careful constructions of the Maslov index and tripleindex in a symplectic Hilbert space. As a byproduct we give intuitively appealingproofs of two theorems of Nicolaescu on the spectral flow of Dirac operators.

As an application of our methods, we carry out a detailed analysis of the η–invariant of the odd signature operator coupled to a flat connection using adiabaticmethods. This is used to extend the definition of the Atiyah–Patodi–Singer ρ–invariantto manifolds with boundary. We derive a “non–additivity” formula for the Atiyah–Patodi–Singer ρ–invariant and relate it to Wall’s non-additivity formula for the sig-nature of even–dimensional manifolds.

Contents

1. Introduction 22. Dirac operators on manifolds with boundary and the self–adjoint Fredholm

Grassmannian 43. The η–invariant and spectral flow 114. The Scott–Wojciechowski theorem 135. Splittings of manifolds and the η–invariant I 176. Maslov index and winding number 257. Splittings of manifolds and the η–invariant II 358. Adiabatic stretching and applications to the Atiyah-Patodi-Singer ρ-invariant 39References 64

The first named author gratefully acknowledges the support of the National Science Foundationunder grant no. DMS-9971020. The second named author was supported by a Heisenberg fellowshipof Deutsche Forschungsgemeinschaft and by the National Science Foundation under grant no. DMS-0072551 .

1

2 PAUL KIRK AND MATTHIAS LESCH

1. Introduction

An intriguing feature of certain spectral invariants is that they behave nicely withrespect to cutting and pasting. Such a feature has several advantages, in particular withrespect to computations. For example, the index of a Dirac operator behaves additivelywith respect to gluing of manifolds. This is not surprising due to the locality of theindex. For higher spectral invariants (e.g. analytic torsion and the η–invariant) cuttingand pasting properties came as a surprise and proofs are nontrivial. The gluing formulafor the η–invariant has a long history (cf. [7] for a historical account). Basically, thereare four different types of proof due to Bunke [9], Wojciechowski [33, 34], Muller [25]and Bruning and Lesch [7]. Bunke’s argument was simplified and generalized by Daiand Freed [12].

While the articles [33, 34, 25, 12] contain proofs of the gluing formula only inR/Z, the original formula of Bunke [9] offers a formula for the integer contributionin terms of indices of certain projections. Unfortunately, these projections are notintrinsically defined and therefore Bunke’s formula is difficult to work with. In [7] it isshown (though not explicitly stated) that the integer contribution can be expressed asthe spectral flow of a naturally defined family of self–adjoint operators.

In the current paper we present another formula for the integer contribution interms of Calderon projectors. This is very satisfactory from a theoretical point of viewsince all ingredients of the formula are defined intrinsically. Moreover, using adiabatictechniques our formula can be made rather explicit; we carry out a detailed analysisfor the odd signature operator.

Given an appropriate orthogonal projection P in the Hilbert space of sections overthe boundary, the domain of a Dirac operator D can be restricted to those sectionswhose restriction to the boundary lie in the kernel of P . Denote the resulting operatorDP . The self–adjoint Fredholm Grassmannian Gr(A) (see Definition 2.1) consists ofthose projections P so thatDP is a self–adjoint discrete Fredholm operator. It contains adistinguished element, namely the Calderon projector for the Dirac operator D. Denoteby η the reduced η–invariant, η(D) = (η(D) + dimkerD)/2. Our main result is thefollowing. (See Theorem 5.9, Theorem 7.4, and Lemma 5.1.)

Theorem. Let D be a Dirac operator on the closed manifold M and let N ⊂ Msplit M into M+ and M−. Assume that D is in product form D = γ( d

dx+A) in a collar

of N , with A self–adjoint. Let P ∈ Gr(A) and let Pt be a smooth path in Gr(A) fromP to the Calderon projector PM+ for D acting on M+. Then

η(D,M) = η(DP ,M+) + η(DI−P ,M

−) + SF(DPt,M+)t∈[0,1] + SF(DI−Pt

,M−)t∈[0,1]= η(DP ,M

+) + η(DI−P ,M−)− τµ(I − PM−, P, PM+).

In particular, taking P = PM+,

η(D,M) = η(DPM+,M+) + η(DI−P

M+,M−).

In these formulas SF denotes the spectral flow, and τµ refers to a Maslov triple indexwe define for appropriate triples of projections. We also prove a more general formula,Theorem 5.10, which holds for any boundary conditions (P,Q), rather than the specialcase (P, I − P ).

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 3

It is well–known that spectral flow and η–invariants are intimately related. It istherefore an interesting feature of our approach that it can be used to give new andconceptually simple proofs of Nicolaescu’s formulas for the spectral flow of a family ofDirac operators [26]. (See Theorems 7.5 and 7.6.)

For purposes of computation it is usually convenient to use the positive spec-tral projection of the tangential operator, P+, rather than the Calderon projec-tor as boundary conditions. According to our theorem this requires computingSF(DPt

,M+)t∈[0,1] +SF(DI−Pt,M−)t∈[0,1] where Pt is a path starting at P+ and ending

at the Calderon projector. In favorable circumstances, such a path (actually its reverse)is obtained by stretching the collar neighborhood of the separating hypersurface. Moreprecisely, replacing M+ by M+ ∪ (N × [−r, 0]) gives a continuous path (as r → ∞)of projections starting at the Calderon projector and limiting essentially to P+. Thisgives a method to obtain computationally useful splitting formulas, and sheds light onthe mechanism of adiabatic stretching.

We carry out this analysis in detail in Section 8 for the odd signature operator.Given a flat connection with holonomy α over an odd–dimensional manifold, we take Dto be the odd signature operator in the corresponding flat bundle. The adiabatic limitof the Calderon projectors for D as the collar is stretched is identified in Theorem 8.5.We use this identification along with the topological invariance of the kernel of D toestablish the formula (cf. (8.32)):

η(D,M) = η(DP+(V+,α),M+) + η(DP−(V−,α),M

−) +m(V+,α, V−,α, α, g).

In this expression V±,α = imH∗(M±;Cnα) → H∗(N ;Cn

α), and m(V+,α, V−,α, α, g) isa real–valued symplectic invariant which depends only on the subspaces V±,α ⊂H∗(N ;Cn

α) and a choice of Riemannian metric on the separating hypersurface N .The projections P±(V±,α) are the sum of the positive/negative spectral projectionsof the tangential operator and the finite–dimensional projection to V±,α. In particularif H∗(N ;Cn

α) = 0 the formula simplifies to

η(D,M) = η(DP+,M+) + η(DP−,M−).

These formulae motivate a definition for the ρ–invariant of a manifold with bound-ary, ρ(X,α, g) (Definition 8.17), which is shown to depend only on the smooth structureof X , the conjugacy class of the representation α, and the choice of Riemannian met-ric g on ∂X . We then prove the following theorem, and discuss its relation to Wall’snon–additivity theorem [32] for the signature of even–dimensional manifolds.

Theorem 8.18. Suppose the closed, odd–dimensional manifoldM contains a hyper-surface N separating M into M+ and M−. Fix a Riemannian metric g on N . Supposethat α : π1(M) → U(n) is a representation, and let τ : π1(M) → U(n) denote the trivialrepresentation. Then

ρ(M,α) = ρ(M+, α, g) + ρ(M−, α, g) +m(V+,α, V−,α, α, g)−m(V+,τ , V−,τ , α, g).

The paper is organized as follows:In Section 2 we review the basic facts about Dirac operators on manifolds with

boundary and the Grassmannian of their boundary value problems.

4 PAUL KIRK AND MATTHIAS LESCH

In Section 3 we introduce the η–invariant and review its basic features. Usingthe Scott–Wojciechowski Theorem [30] we prove in Section 4 a formula describing thedependence on the choice of boundary condition of the η–invariant of a Dirac operatoron a manifold with boundary (Theorem 4.4).

Section 5 deals with splittings of manifolds. We prove a result on the behavior of thespectral flow under splittings (Corollary 5.6) and the gluing formula for the η–invariant(Theorem 5.10).

Section 6 contains careful constructions of various forms of the Maslov index forfamilies of self–adjoint projections in a Hermitian symplectic Hilbert space. Conventionsmust be set to deal with degenerate situations when defining symplectic invariants, andwe carefully construct the various invariants consistently and in such a way that theymatch our choice of convention for the spectral flow.

A byproduct of our considerations are new proofs of (generalizations of) two the-orems by Nicolaescu [26] identifying the spectral flow of a family of Dirac operatorswith a Maslov index involving the Calderon projectors and boundary conditions. Theseresults (Theorem 7.5 for manifolds with boundary and Theorem 7.6 for split manifolds),together with an improvement (Theorem 7.7) of our gluing formula for the η–invariantwhich allows more general boundary conditions, are presented in Section 7.

Finally, in Section 8 we apply our splitting results for the η–invariant to the specialcase of the odd signature operator coupled to a flat connection. By making use of themethod of adiabatic stretching of the collar of a separating hypersurface and the factthat the dimension of the kernels of these operators are topological, i.e. independentof the Riemannian metric, we obtain a splitting formula for the Atiyah–Patodi–Singerρα invariant. The main tool introduced in this section is Theorem 8.5, which gives aprecise identification of the adiabatic limit of the Calderon projectors in this setting.We end the paper with an examination of the role adiabatic stretching plays in additionformulas for the η–invariants of general Dirac operators.

2. Dirac operators on manifolds with boundary and the self–adjoint

Fredholm Grassmannian

We begin by describing the set–up of Dirac operators on a manifold with boundary.Let X denote a compact smooth Riemannian manifold with boundary ∂X . We

fix an identification of a neighborhood of ∂X in X with ∂X × [0, ε). Let E → Xbe a complex Hermitian vector bundle and suppose that D : C∞(E) → C∞(E) isa symmetric Dirac operator, i.e. a symmetric first–order operator whose square is ageneralized Laplacian (the square of the leading symbol of D is scalar and given by themetric tensor). The symmetry is measured with respect to the L2 inner product; thuswe assume that if φ1, φ2 ∈ C∞(E) are supported in the interior of X then

X

〈Dφ1, φ2〉Exdx =

X

〈φ1, Dφ2〉Exdx.

A Dirac operator satisfies the unique continuation property [5].In this paper we will deal only with the product case, i.e. we assume that the

restriction of D to the collar takes the form D = γ( ddx

+ A), where γ : E|∂X → E|∂X

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 5

is a bundle endomorphism and A : C∞(E|∂X) → C∞(E|∂X) is a first–order self–adjointelliptic differential operator on the closed manifold ∂X (called the tangential operator)satisfying

γ2 = −I, γ∗ = −γ, and γA = −Aγ. (2.1)

Note that A is assumed to be independent of x for x ∈ [0, ε).The operator D : C∞(E) → C∞(E) can be extended to an unbounded self–adjoint

operator on L2(E) by imposing appropriate boundary conditions. Since D is a firstorder operator, it can be extended to a bounded operator H1(E) → L2(E), whereHs(E) denotes the Sobolev space of sections of E with s derivatives in L2. Givenan orthogonal projection P : L2(E|∂X) → L2(E|∂X) define DP to be D acting on thedomain

D(DP ) :=φ ∈ L2(E)

∣∣ φ ∈ H1(E) and P (φ|∂X) = 0⊂ L2(E).

We will consider the operators DP for a certain class of projections P which we nowintroduce. Let

P>0 : L2(E|∂X) −→ L2(E|∂X)

denote the positive spectral projection for the self–adjoint tangential operator A :C∞(E|∂X) → C∞(E|∂X); thus if ψλ is a basis of L2(E|∂X) with Aψλ = λψλ, thenP>0(

∑aλψλ) =

∑λ>0 aλψλ.

Definition 2.1. Define the self–adjoint Fredholm Grassmannian Gr(A) to be theset of maps P : L2(E|∂X) → L2(E|∂X) so that

(1) P is pseudo–differential of order 0,(2) P = P ∗, P 2 = P , i.e. P is an orthogonal projection,(3) γPγ∗ = I − P ,(4) (P>0, P ) form a Fredholm pair, that is,

P>0| imP : imP → imP>0

is Fredholm.

The Grassmannian Gr(A) is topologized using the norm topology on bounded op-erators.

Remark 2.2.

1. We note that a P ∈ Gr(A) also acts as a (non–orthogonal) projection in theSobolev space Hs(E) for all s ∈ R. This follows from (1).

2. We obtain the same Grassmannian if we replace P>0 in (4) by any pseudo–differential orthogonal projection Q such that P>0−Q is smoothing. This followsimmediately from the following general fact:

Let P,Q,R be orthogonal projections in the Hilbert space H suchthat Q−R is compact. Then (P,Q) is a Fredholm pair if and only if(P,R) is a Fredholm pair.

6 PAUL KIRK AND MATTHIAS LESCH

This fact can be seen as follows: by [3, Prop. 3.1] (P,Q) is Fredholm if andonly if ±1 6∈ specess(P − Q). Since Q − R is compact this is equivalent to±1 6∈ specess(P − R). Applying again [3, Prop. 3.1] the latter is the case if andonly if (P,R) is Fredholm.

If P ∈ Gr(A), then DP is self–adjoint, Fredholm, and has compact resolvent; inparticular its spectrum is discrete and each eigenvalue has finite multiplicity. Thesefacts follow since (D,P ) is a well–posed boundary value problem in the sense of R. T.Seeley [31]. A general reference for boundary value problems for Dirac type operatorsis the monograph [5]. A different approach is presented in [8, 6].

It will be necessary to consider a more restricted class of projections, those thatdiffer from P>0 by a smoothing operator. Define Gr∞(A) ⊂ Gr(A) by

Gr∞(A) =P ∈ Gr(A)

∣∣ P − P>0 is a smoothing operator. (2.2)

Again, in (2.2) we can replace P>0 by any pseudo–differential orthogonal projection Qsuch that P>0 −Q is smoothing.

The projection P>0 does not lie in Gr(A) unless kerA = 0, since the third conditiondoes not hold for P = P>0 if kerA 6= 0. It is convenient to specify a finite rankperturbation of P>0 which does lie in Gr(A).

Notice that γ leaves kerA invariant. It is well–known that since (∂X,A) “bounds”(X,D), the i and −i eigenspaces of γ acting on kerA have the same dimension [27,Chap. XVII]. This implies that there are subspaces L ⊂ kerA satisfying γ(L) =L⊥ ∩ kerA (such subspaces are called Lagrangian subspaces; see Definition 2.8 below).Given a Lagrangian subspace L ⊂ kerA define

P+(L) = projL + P>0. (2.3)

Then P+(L) differs from P>0 by the projection onto L, a subspace of kerA, which con-sists only of smooth functions. Since P>0 is a 0th order pseudo–differential projection,so is P+(L). It is straightforward to check that P+(L) ∈ Gr∞(A).

We call P+(L) the Atiyah–Patodi–Singer projection corresponding to the La-grangian subspace L. Notice that P+(L) depends only on the tangential operatorA and the choice of L; in particular it is unchanged if D is altered in the interior of X .

There is a canonical projection in Gr(A) determined by the operator D which willplay a special role in what follows, namely the Calderon projector PX . It is defined asthe orthogonal projection onto the Cauchy data space

LX := r(kerD : H1/2(E) −→ H−1/2(E)

)⊂ L2(E|∂X). (2.4)

Here r denotes the restriction to the boundary. The trace operator r is a priori onlydefined on Hs(E) for s > 1/2 but one can show that r defines a bounded map from theH1/2–kernel of D into L2(E|∂X) (see [5] for a proof).

The Calderon projector PX = projLXlies in Gr∞(A) [29, Prop. 2.2], [17, Prop.

4.1]. The unique continuation property for D implies that

r :(kerD : H1/2(E) −→ H−1/2(E)

)−→ L2(E|∂X)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 7

is injective, so that to any vector x in the image of PX we can assign a unique solutionto Dφ = 0 on X with φ ∈ H1/2 and r(φ) = x. This makes it possible to identify thekernel of D with boundary condition given by a projection P and the intersection ofthe Cauchy data space with the kernel of P , as in the following lemma.

Lemma 2.3. Let P ∈ Gr(A). Then

kerP| imPX= imPX ∩ kerP = γ(kerPX) ∩ kerP

and this space is isomorphic to the kernel of DP . Thus DP is invertible if and only ifimPX ∩ kerP = 0. In particular DPX

is invertible.

Proof. If φ ∈ kerD, then by definition the restriction of φ to the boundary of Xlies in the image of the Calderon projector PX . In particular, if φ ∈ kerDP , then therestriction of φ to the boundary lies in the intersection of kerP and the image of PX .The unique continuation property for D implies that this intersection is exactly thekernel of DP , i.e. the kernel of DP is isomorphic to kerP| imPX

.As a discrete self–adjoint operator, DP is invertible if and only if kerDP = 0.

Moreover, PX is a self–adjoint projection satisfying the equation γPXγ = −(I − PX).Thus imPX = ker(I − PX) = γ(kerPX).

In a rough sense the Atiyah–Patodi–Singer projection P+(L) and the Calderonprojector PX are opposites: P+(L) is determined entirely by the boundary data, i.e.the tangential operator A acting on ∂X (and the choice of L), whereas PX depends onall of X and D.

For future reference we note the following special case of a result due to K. Woj-ciechowski.

Proposition 2.4. The Grassmannians Gr(A),Gr∞(A) are path connected. For afixed P ∈ Gr∞(A) (resp. Gr(A)) the space Q ∈ Gr∞(A) | kerQ ∩ imP = 0 (resp.Q ∈ Gr(A) | kerQ ∩ imP = 0) is path connected.

Remark 2.5. This result could also be proved using Proposition 6.5 below (resp.its analog for pseudo–differential Grassmannians) and properties of the unitary group.

Proof. The first statement is a special case of [15, Appendix B], where the ho-motopy groups of Gr∞(A) and Gr(A) are computed. The path connectedness ofQ ∈ Gr∞(A)

∣∣ kerQ ∩ imP = 0was proved in Proposition 5.1 of [30]. The path

connectedness ofQ ∈ Gr(A)

∣∣ kerQ∩ imP = 0can be proved along the same lines:

if kerQ ∩ imP = 0 then ‖Q − P‖ < 1 and hence Qs := ZsPZ−1s , 0 ≤ s ≤ 1, where

Zs := I + s(Q− P )(2P − I), is a path in Gr(A) connecting P and Q (cf. e.g. [6, Sec.3]).

Notice that Gr(A) and Gr∞(A) can also be defined by replacing P>0 by P+(L) orPX in the fourth condition defining Gr(A), and in (2.2).

8 PAUL KIRK AND MATTHIAS LESCH

We next discuss two alternative perspectives on the Grassmannian Gr(A), identify-ing this space with the space of certain unitary operators on a Hilbert space, and alsowith certain Lagrangian subspaces of a symplectic Hilbert space.

The bundle endomorphism γ : E|∂X → E|∂X induces a decomposition of E|∂X =Ei⊕E−i into the ±i eigenbundles and consequently we get a decomposition of L2(E|∂X)into the ±i eigenspaces,

L2(E|∂X) = L2(Ei)⊕ L2(E−i) =: Ei ⊕ E−i. (2.5)

Given P ∈ Gr(A), write

P =1

2

(A BC D

)

with respect to the decomposition (2.5). Then P = P ∗ implies C = B∗. The conditionsγPγ∗ = I − P and γ∗ = −γ imply that A = D = I, and the condition P 2 = P impliesthat BB∗ = I = B∗B. This proves the first part of the following lemma.

Lemma 2.6. If P ∈ Gr(A), then with respect to the decomposition (2.5), P can bewritten in the form

P =1

2

(I T ∗

T I

), (2.6)

where T is a 0th order pseudo–differential isometry from Ei onto E−i. Conversely, givensuch an isometry T , then

1

2

(I T ∗

T I

)

is a pseudo–differential projection satisfying (1), (2), (3) of Definition 2.1.Given

P =1

2

(I T ∗

T I

), Q =

1

2

(I S∗

S I

),

satisfying (1), (2), (3) of Definition 2.1, then:

(1) (P,Q) form a Fredholm pair if and only if −1 6∈ specess T∗S,

(2) (P,Q) is invertible if and only if −1 6∈ spec T ∗S,(3) kerP ∩ imQ is canonically isomorphic to ker(I + T ∗S),(4) P −Q is smoothing if and only if T ∗S − I is smoothing.

In particular, if Q = P+(L) for some Lagrangian L ⊂ kerA, then P ∈ Gr(A) if andonly if −1 6∈ specess T

∗S.

Proof. The first part was proved above. Since S∗S = I = SS∗, any element in

L2(E|∂X) = Ei ⊕ E−i can be written in the form

(xSy

)for x, y ∈ Ei. Since

Q

(xSy

)=

1

2

(x+ y

S(x+ y)

),

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 9

it follows that imQ =( x

Sx

) ∣∣ x ∈ Ei

. Thus the restriction of P to the image of Q

is

P

(xSx

)=

1

2

((I + T ∗S)xT (I + T ∗S)x

).

It follows that (P,Q) is Fredholm (i.e. P restricts to a Fredholm operator on theimage of Q) if and only if I + T ∗S is Fredholm, which occurs precisely when −1 is notin the essential spectrum of T ∗S. Similarly (P,Q) is invertible (i.e. the restriction ofP to the image of Q defines an isomorphism onto the image of P ) if and only if −1 isnot in the spectrum of T ∗S. The same argument also shows (3).

Finally, since

P −Q =

(0 T ∗ − S∗

T − S 0

),

P −Q is smoothing if and only if T −S is smoothing. Here we use that the projections12i(i±γ) onto E±i are differential operators of order 0. Since T, S are pseudo–differential

and unitary the operator T − S is smoothing if and only if T ∗S − I is smoothing.

Let U (Ei, E−i) denote the set of unitary isomorphisms from Ei to E−i. Then P 7→ Tdefines a map

Φ : Gr(A) −→ U (Ei, E−i), (2.7)

i.e.,

P =1

2

(I Φ(P )∗

Φ(P ) I

).

More abstractly, consider the group U of unitary pseudo–differential isomorphismsEi → Ei. Let

UFred =U ∈ U

∣∣ −1 6∈ specess U, (2.8)

and

U∞ =U ∈ UFred

∣∣ U − I is a smoothing operator. (2.9)

Then given any P ∈ Gr∞(A), the map

U 7→1

2

(I (Φ(P )U)∗

Φ(P )U I

)(2.10)

defines homeomorphisms

UFred −→ Gr(A)

and

U∞ −→ Gr∞(A).

Another useful description of Gr(A) and Gr∞(A) is in terms of Lagrangian sub-spaces.

10 PAUL KIRK AND MATTHIAS LESCH

Lemma 2.7. Let (H, 〈 , 〉) be a separable complex Hilbert space and γ : H → H anisomorphism satisfying γ2 = −I, γ∗ = −γ. Then there exists a subspace L ⊂ H suchthat γ(L) = L⊥ if and only if dimker(γ − i) = dimker(γ + i).

Proof. Suppose L ⊂ H is a subspace with γ(L) = L⊥. Then it is easy to checkthat the orthogonal projections π± : L → ker(γ ± i) are isomorphisms and hencedim ker(γ + i) = dimker(γ − i).

Conversely, if dim ker(γ+ i) = dimker(γ− i) then let T : ker(γ− i) → ker(γ+ i) be

a unitary isomorphism. Then L =( x

Tx

) ∣∣ x ∈ ker(γ − i)is a subspace satisfying

γ(L) = L⊥.

Definition 2.8. A Hermitian symplectic Hilbert space is a separable complexHilbert space together with an isomorphism γ : H → H satisfying γ2 = −I, γ∗ = −γand such that the i and −i eigenspaces of γ have the same dimension (i.e. if H isinfinite–dimensional we require that both eigenspaces are infinite–dimensional). Thesymplectic form is the skew-Hermitian form

ω(x, y) := 〈x, γy〉. (2.11)

A Lagrangian subspace L ⊂ H of a Hermitian symplectic Hilbert space is a subspaceso that γ(L) = L⊥. A Lagrangian subspace is automatically closed.

The space L2(E|∂X) together with the map γ is a Hermitian symplectic Hilbertspace. The space kerA is a finite–dimensional Hermitian symplectic Hilbert space since(∂X,A) bounds (X,D).

Given P ∈ Gr(A), the kernel of P is a Lagrangian subspace, since kerP is orthogonalto γ(kerP ). Notice that the kernel of P can be expressed as the graph of −Φ(P ),

kerP =( x

−Φ(P )x

) ∣∣ x ∈ Ei

⊂ L2(E|∂X).

This gives a third characterization of Gr(A) as follows. We define L to be the set ofLagrangian subspaces of L2(E|∂X) whose associated projections are pseudo–differentialof order 0. The Cauchy data space, LX , (the image of the Calderon projector) is aLagrangian subspace of L2(E|∂X).

Define

LFred =L ∈ L

∣∣ (L, γ(LX)) is a Fredholm pair of subspaces, (2.12)

and

L∞ =L ∈ LFred

∣∣ projL − projLXis a smoothing operator

. (2.13)

Then we have homeomorphisms

LFred −→ Gr(A)

and

L∞ −→ Gr∞(A).

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 11

The identifications of L , Gr(A), and U are determined by the conditions that

L ∈ LFred, P ∈ Gr(A), and T ∈ UFred(Ei, E−i)

correspond ifL = imP = graph of T and T = Φ(P ).

3. The η–invariant and spectral flow

It was mentioned in the last section that DP is the self–adjoint realization of a well–posed boundary value problem and hence it is a discrete operator in the Hilbert spaceL2(E). For the discussion of ζ– and η–functions we need the more refined analysis ofthe heat trace of DP . The ζ– and η–functions of DP are defined, for Re(s) >> 0, by

η(DP ; s) := tr(DP |DP |−s−1) =

λ∈specDP \0

sign(λ)|λ|−s,

ζ(DP ; s) := tr(D−sP ) =

λ∈specDP \0

λ−s

=1

2

(ζ(D2

P ; s/2) + η(DP ; s))+ e−iπs1

2

(ζ(D2

P ; s/2)− η(DP ; s)).

(3.1)

Theorem 3.1. For P ∈ Gr(A) the functions ζ(DP ; s), η(DP ; s) extend meromor-phically to the whole complex plane with poles of order at most 2. If P ∈ Gr∞(A)then η(DP ; s) and ζ(DP ; s) are regular at s = 0. Moreover ζ(DP ; 0) is independent ofP ∈ Gr∞(A).

That the ζ– and η–functions extend meromorphically has been proved in increasinggenerality in [15], [18], [19], [7], [35], and [17]. The definitive treatment of all well–posed boundary value problems is given in [17]. The proof of the statement aboutregularity at s = 0 can be found in [35]. The methods of [17] show that the assumptionP ∈ Gr∞(A) can be somewhat relaxed [16]. Finally, that ζ(DP ; 0) is independent ofP ∈ Gr∞(A) is [35, Prop. 0.5].

Definition 3.2. The η-invariant of DP , η(DP ), is defined to be the constant termin the Laurent expansion of η(DP ; s) at s = 0, i.e.

η(DP ; s) = as−2 + bs−1 + η(DP ) +O(s).

We also give a symbol to a convenient normalization of the η–invariant.

Definition 3.3. The reduced η–invariant is defined to be

η(DP ) = (η(DP ) + dimkerDP ) /2. (3.2)

We continue with a discussion of the spectral flow and its relation to the η–invariant.Suppose one is given a smooth path of Dirac type operators Dt : C

∞(E) → C∞(E), t ∈[0, 1], over X so thatDt = γ( d

dx+At) on the collar. Choose a smooth path of projections

Pt so that Pt ∈ Gr(At) for t ∈ [0, 1]. Then the family DP (t) := (Dt)Ptis in particular

a graph continuous family of self–adjoint discrete operators. As a consequence, theeigenvalues of DPt

vary continuously (as a general reference see [21]). The spectral

12 PAUL KIRK AND MATTHIAS LESCH

flow of the family DP (t), which we denote by SF(DP (t))t∈[0,1] or just SF(DP (t)), is theinteger defined (roughly) to be the difference in the number of eigenvalues that startnegative and end non-negative and the number of eigenvalues that start non-negativeand end negative (see [5, 11] for a precise definition). Notice that we have chosen aparticular convention for dealing with zero eigenvalues. This convention is often calledthe (−ε,−ε) spectral flow in the literature, since it corresponds to intersecting thegraph of the spectrum as a function of t with the line from (0,−ε) to (1,−ε).

The 1–parameter family of η–invariants η(DP (t)) ∈ R will in general not varysmoothly with respect to t ∈ [0, 1]. However, it follows from the work of G. Grubb [17]that the reduction modulo integers of the reduced η–invariant η(DP (t)) varies smoothlywith t. In particular, the real valued function

u 7→

∫ u

0

d

dt

(η(DP (t))

)dt

is smooth.In general, given a smooth function f : [0, 1] → R/Z = S1, the expression u 7→

c +∫ u

0dfdtdt is just an explicit formula for the unique smooth lift of f to the universal

cover R of S1 starting at c ∈ R. Thus if f and g are (possibly discontinuous) functionsfrom [0, 1] to R so that the reductions of f and g modulo Z are smooth and agree, thenthe smooth real–valued functions u 7→

∫ u

0dfdtdt and u 7→

∫ u

0dgdtdt coincide.

Lemma 3.4. Suppose that Dt, t ∈ [0, 1], is a smooth path of symmetric Dirac typeoperators as above, and Pt ∈ Gr(At) is a smooth path, giving a smooth path of self–adjoint discrete operators DP (t).

Then

η(DP (1))− η(DP (0)) = SF(DP (t))t∈[0,1] +1

2

∫ 1

0

d

dt(η(DP (t)))dt.

Moreover, if the dimension of the kernel of DP (t) is independent of t, then the functiont 7→ η(DP (t)) is smooth.

Proof. We only sketch the proof, since this fact is well–known, at least when theη–function is regular at s = 0, and the general case is proven by the same argument,because the pole of the η–function at s = 0 is determined by the asymptotics of thespectrum, whereas the spectral flow depends only on the small eigenvalues.

Given r ∈ [0, 1], choose an ε > 0 so that ±ε does not lie in the spectrum of DP (r).Applying standard results from perturbation theory [21] we infer that ±ε does not liein the spectrum of DP (t) for t close enough to r, say t ∈ [t0, t1]. Moreover the spanof those eigenvectors of DP (t) whose eigenvalues lie in (−ε, ε) varies continuously fort ∈ [t0, t1].

Thus we can write η(DP (t); s) for t ∈ [t0, t1] and Re(s) >> 0 as a sum

η(DP (t); s) =∑

λ∈specDP (t),0<|λ|<ε

sign(λ)|λ|−s +∑

λ∈specDP (t),|λ|>ε

sign(λ)|λ|−s

= η<ε(DP (t); s) + η>ε(DP (t); s).

(3.3)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 13

The sum η<ε(DP (t); s) is finite, and so its analytic continuation to s = 0 is integervalued:

η<ε(DP (t); 0) =∑

λ∈specDP (t),0<|λ|<ε

sign(λ).

Thus

η<ε(DP (t1); 0)− η<ε(DP (t0); 0)

= 2 SF(DP (t))t∈[t0,t1] + dimkerDP (t0)− dimkerDP (t1).(3.4)

Notice that this equation depends on our choice of convention for defining the spectralflow. The function η>ε(DP (t); s) = η(DP (t); s) − η<ε(DP (t); s) varies smoothly in t ∈[t0, t1] since we have subtracted the eigenvalues that cross zero, and since no eigenvaluesequal ±ε in this interval. If we define η>ε(DP (t)) similarly to Definition 3.2 thenη>ε(DP (t)) is smooth and

η>ε(DP (t)) ≡ η(DP (t))modZ.

Therefore, using (3.4) we obtain∫ t1

t0

d

dt

(η(DP (t))

)dt

=

∫ t1

t0

d

dt

(η>ε(DP (t))

)dt

= η>ε(DP (t1))− η>ε(DP (t0))

= η(DP (t1))− η<ε(DP (t1))− η(DP (t0)) + η<ε(DP (t0))

= η(DP (t1))− η(DP (t0))− 2 SF(DP (t))t∈[t0,t1]

+ dimkerDP (t1)− dimkerDP (t0).

(3.5)

Dividing by 2 proves the lemma over the interval [t0, t1]. The general case is obtainedby covering the interval [0, 1] by small subintervals and adding the results.

For the last assertion, notice that if the dimension of kerDP (t) is independent of t,then SF(DP (t))t∈[0,s] = 0 for all s ∈ [0, 1].

4. The Scott–Wojciechowski theorem

The theorem of Scott and Wojciechowski [30] identifies the regularized ζ–deter-minant of a boundary–value problem for a Dirac operator with a Fredholm determinantof the associated boundary projection. In this section we summarize and slightly extendthat part of their result which we need, in the language of this article. Briefly, theirtheorem shows that the reduction mod Z of the η–invariant of DP for P ∈ Gr∞(A) andthe Fredholm determinant of the unitary map Φ(P ) which corresponds to P via (2.10)agree up to a constant independent of P . The important consequence for this articleis that the modZ reduction of the η–invariant for a manifold with boundary dependsonly on the boundary data and the Calderon projector.

In this section D denotes a fixed Dirac type operator on a manifoldX with boundary∂X and A denotes its tangential operator.

14 PAUL KIRK AND MATTHIAS LESCH

Before stating the Scott–Wojciechowski theorem, let us briefly recall the ζ–regularized determinant. Let P ∈ Gr∞(A). Then ζ(DP ; s) is regular at s = 0 andone puts

detζ DP :=

exp

(− ζ ′(DP ; 0)

), 0 6∈ specDP ,

0, 0 ∈ specDP .(4.1)

In view of (3.1) and Theorem 3.1 a straightforward calculation shows for DP invertible

detζ DP = exp(iπ

2

(ζ(D2

P ; 0)− η(DP ))−

1

2ζ ′(D2

P ; 0)). (4.2)

We emphasize that the regularity of η(DP ; s) and ζ(DP ; s) at s = 0 is essential for(4.2) to hold. (4.2) implies that in general (detζ D)2 6= detζ(D

2). Note that Fredholmdeterminants are multiplicative, i.e. if S, T are operators of determinant class in aHilbert space then detF(ST ) = detF(S) detF(T ), where detF denotes the Fredholmdeterminant.

With these preparations, the Scott–Wojciechowski theorem reads as follows.

Theorem 4.1. Let P ∈ Gr∞(A). Then

detζ(DP ) = detζ(DPX) detF

(I + Φ(PX)Φ(P )∗

2

). (4.3)

This result was proved for M odd–dimensional in [30, Thms. 0.1, 1.4]. An alterna-tive proof which applies to all dimensions and to slightly more general operators willbe presented in [23].

In view of (4.2) the Scott–Wojciechowski theorem can be applied to express thedependence of η(DP ) on P in terms of Fredholm determinants.

Let P,Q ∈ Gr∞(A). Since Φ(Q)Φ(P )∗ − I is a smoothing operator, it is of traceclass and hence

I + Φ(Q)Φ(P )∗

2= I +

Φ(Q)Φ(P )∗ − I

2(4.4)

is of determinant class. In particular, I+Φ(PX)Φ(P )∗

2is of determinant class and thus the

right hand side in (4.3) is well–defined.Also, Φ(P )Φ(Q)∗ is of determinant class. Hence the determinant detF(Φ(P )Φ(Q)

∗)is defined and lies in U(1) since Φ(P )Φ(Q)∗ is unitary.

Theorem 4.2. Let P,Q ∈ Gr∞(A). Then

e2πi(η(DP )−η(DQ)) = detF(Φ(P )Φ(Q)∗). (4.5)

Proof. Assume first that P is the Calderon projector PX and that the pair (PX , Q)is invertible. By Lemma 2.3 this means that DPX

and DQ are invertible. PuttingTheorem 4.1 and (4.2) together and taking into account that ζ(D2

P ; 0) is independentof P (Theorem 3.1), we obtain

eiπ2

(η(DPX

)−η(DQ))e

12

(ζ′(D2

PX;0)−ζ′(D2

Q;0))=

detζ(DQ)

detζ(DPX)= detF

(I + Φ(PX)Φ(Q)∗

2

), (4.6)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 15

and thus

detF( I+Φ(PX)Φ(Q)∗

2

)∣∣detF

( I+Φ(PX)Φ(Q)∗

2

)∣∣ = eiπ2

(η(DPX

)−η(DQ)). (4.7)

Since Φ(PX)Φ(Q)∗ − I is of trace class we may choose a self–adjoint trace class

operator H such that eiH = Φ(PX)Φ(Q)∗. Then

detF(I + Φ(PX)Φ(Q)

2

)2= detF

(I + eiH

2

)2

= detF(eiH cosh2(H/2)

)

= detF(Φ(PX)Φ(Q)

∗)detF

(cosh2(H/2)

),

(4.8)

where we have used the multiplicativity of the Fredholm determinant in the last line.Consequently

detF( I+Φ(PX)Φ(Q)∗

2

)2∣∣detF

( I+Φ(PX)Φ(Q)∗

2

)∣∣2 = detF(Φ(PX)Φ(Q)

∗). (4.9)

Putting together (4.7) and (4.9) we obtain (4.5) for P = PX and Q ∈ Gr∞(A) such that(PX , Q) is an invertible pair. However, since both sides of (4.5) depend continuouslyon Q, (4.5) remains valid for all Q ∈ Gr∞(A). Finally, if P,Q ∈ Gr∞(A) are arbitrarythen

e2πi(η(DP )−η(DQ)) = e2πi(η(DP )−η(DPX))e2πi(η(DPX

)−η(DQ))

= detF(Φ(P )Φ(PX)

∗)detF

(Φ(PX)Φ(Q)

∗)

= detF(Φ(P )Φ(Q)∗

).

(4.10)

We will use the following convenient form of the Scott–Wojciechowski theorem. Weconsider the reals R as the universal cover of U(1) via the map r 7→ e2πir.

Corollary 4.3. Let Pt, t ∈ [0, 1], be a smooth path in Gr∞(A). Then the map

s 7→1

2

∫ s

0

d

dt(η(DPt

))dt

is the unique lift to R of the map [0, 1] → U(1)

s 7→ detF(Φ(Ps)Φ(P0)

∗).

In preparation for the next theorem, suppose that P ∈ Gr∞(A). From Lemma 2.3we know that DP is invertible if and only if kerPX ∩γ(kerP ) = 0 where PX denotes theCalderon projector; by Lemma 2.6 this happens if and only if −1 6∈ spec(Φ(P )Φ(PX)

∗).In fact, the kernel of DP is canonically isomorphic to ker(I + Φ(P )Φ(PX)

∗).Using the functional calculus we can define the operator log(Φ(P )Φ(PX)

∗). Thechoice of the branch of log will be essential in what follows. We define log : C\0 → C

as follows

log(reit) = ln r + it, r > 0,−π < t ≤ π. (4.11)

16 PAUL KIRK AND MATTHIAS LESCH

Since −1 6∈ specess(Φ(P )Φ(PX)∗), −1 is an isolated point in the spectrum of

Φ(P )Φ(PX)∗ and thus we can choose a holomorphic branch of the logarithm which coin-

cides on spec(Φ(P )Φ(PX))∗ with log defined in (4.11). The so defined log(Φ(P )Φ(PX)

∗)is of trace class and

tr log(Φ(P )Φ(PX)∗) ≡ log detF

(Φ(P )Φ(PX)

∗)mod2πiZ.

After these preparations we can improve Theorem 4.2 as follows.

Theorem 4.4. Let X be a compact manifold with boundary and let D be a Diractype operator such that in a collar ∂X × [0, ε) of the boundary D takes the form D =γ( d

dx+A) with A, γ as in (2.1). Let Φ be the map defined in (2.7). Then for P ∈ Gr∞(A)

we have

η(DP )− η(DPX) =

1

2πitr log(Φ(P )Φ(PX)

∗).

Proof. We assume first that DP is invertible. DPXis invertible by Lemma 2.3. In

view of Proposition 2.4 and Lemma 2.3 the space of those P ∈ Gr∞(A) so that DP isinvertible is path connected. Choose a smooth path Pt in Gr∞(A) starting at PX andending at P so that DPt

is invertible for all t.The spectral flow of DPt

equals zero since the kernel is zero along the path and soLemma 3.4 shows that t 7→ η(DPt

) is smooth. Hence

t 7→ η(DPt)− η(DPX

) (4.12)

is smooth. Also, the map

t 7→1

2πitr log(Φ(Pt)Φ(PX)

∗) (4.13)

is smooth since −1 6∈ spec(Φ(Pt)Φ(PX)∗) for all t and hence log is holomorphic on

spec(Φ(Pt)Φ(PX)∗).

Theorem 4.2 states that the two smooth functions of (4.12) and (4.13) are the liftsto R of the same function to U(1) = R/Z, and they both start at 0. Hence they coincidefor all t.

Now let P ∈ Gr∞(A) be arbitrary. We may choose a path (Pt)−ε≤t≤ε in Gr∞(A)such that (Pt, PX) is invertible for t 6= 0, P0 = P , and such that at t = 0 exactlyk = dimkerDP0

eigenvalues of Φ(Pt)Φ(PX)∗ cross −1 from the upper half plane to the

lower half plane and no eigenvalues cross from the lower half plane to the upper halfplane. To see this let R be the orthogonal projection onto ker(I + Φ(P )Φ(PX)

∗). Theprojection R is a pseudo–differential operator. Now put

Φ(Pt) :=(ei(π+t)R ⊕ (I − R)Φ(P )Φ(PX)

∗)Φ(PX). (4.14)

By our choice of log we then have

1

2πitr log(Φ(P0)Φ(PX)

∗) = limt→0−

1

2πitr log(Φ(Pt)Φ(PX)

∗). (4.15)

Moreover, for t 6= 0 we have from the first part of this proof

η(DPt)− η(DPX

) =1

2πitr log(Φ(Pt)Φ(PX)

∗). (4.16)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 17

From Lemma 3.4, (4.14) and (4.16) one infers SF(DPt)−ε≤t≤ε = −k and since

dim kerDP0= k at t = 0 exactly k eigenvalues of DPt

cross 0 from + to − and noeigenvalues cross from − to +. Hence

η(DP0)− η(DPX

) = limt→0−

η(DPt)− η(DPX

)

= limt→0−

1

2πitr log(Φ(Pt)Φ(PX)

∗)

=1

2πitr log(Φ(P0)Φ(PX)

∗),

(4.17)

completing the proof.

5. Splittings of manifolds and the η–invariant I

We consider now the gluing problem for the η–invariant. Suppose we are given aclosed manifold M containing a separating hypersurface N ⊂ M . We consider onlyDirac operators D on M so that in a collar neighborhood [−ε, ε]×N of N , D has theform D = γ( d

dx+ A) as in (2.1).

Let M cut denote the compact manifold with boundary obtained by cuttingM alongN . Thus M cut is the disjoint union of two submanifolds M+ and M−, with ∂M+

and ∂M− canonically identified with N . To apply the results of the previous section,we reparameterize the collar of M− as ∂M− × [0, ε] with x = 0 corresponding to theboundary. See the following figure.

M

M+

N

Mcut

M+

M

M

The manifolds M and M cut

18 PAUL KIRK AND MATTHIAS LESCH

The H1–Sections of a bundle E over M correspond to sections f ∈ H1(E|Mcut)over M cut so that f|∂M+ = f|∂M− with respect to the canonical identifications ∂M± =N . More precisely, the restriction of the section f to the boundary of M cut lies inH1/2(E|∂Mcut) ⊂ L2(E|∂Mcut). The identification of ∂M cut with two copies of N gives acanonical decomposition

L2(E|∂Mcut) = L2(E|N)⊕ L2(E|N) (5.1)

where the first factor corresponds to ∂M+ and the second to ∂M−. The restriction ofa section f over M cut to the boundary can thus be written as (f+, f−), and the sectionsover M correspond exactly to those f so that f+ = f−.

On the collar of M cut, The operator D takes the form

D =

(γ 00 −γ

)(d

dx+

(A 00 −A

))=: γ(

d

dx+ A) (5.2)

with respect to the decomposition (5.1) (this is because of the change of parameteriza-tion of the collar of M−).

Note that the (closed) diagonal subspace

∆ = (f, f) | f ∈ L2(E|N) ⊂ L2(E|N)⊕ L2(E|N)

is Lagrangian. In fact:

1. ∆ is orthogonal to γ(∆) since

〈(f, f), γ(g, g)〉 = 〈f, γg〉+ 〈f,−γg〉 = 0.

2. ∆ + γ(∆) = L2(E|N)⊕ L2(E|N) since

(f, g) =1

2((f + g, f + g) + γ(−γf + γg,−γf + γg)) .

The orthogonal projection to ∆⊥ will be denoted by P∆. It is called the continuoustransmission projection. By construction an H1–section f over M cut defines an H1–section over M if and only if the restriction (f+, f−) of f to the boundary satisfies

P∆(f+, f−) = 0. Note that P∆ ∈ Gr(A) since DP∆is canonically identified with the

(Fredholm) operator D acting on the closed manifold M .With respect to the decomposition (5.1) the operator P∆ takes the form

P∆ =1

2

(1 −1

−1 1

). (5.3)

In a strict sense, the projection P∆ is not in Gr(A) since it does not act as a pseudo–differential operator on E|∂Mcut. Namely, since (5.3) contains off–diagonal terms thetwo copies of of N ⊂ ∂M cut interact and hence P∆ is a Fourier integral operator.However, P∆ is pseudo–differential on the bundle E|N ⊕E|N over N . This is only a mildgeneralization of the situation of Section 2 and we refrain from formalizing it. From

now on Gr(A) is to be understood as the set of those pseudo–differential operators onthe bundle E|N ⊕ E|N over N which satisfy (2), (3), (4) of Definition 2.1. It is fairlyclear that the results of the previous sections also apply to this situation.

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 19

There is a natural map

Gr(A)×Gr(−A) −→ Gr(A), (P,Q) 7→

(P 00 Q

)(5.4)

with respect to the decomposition (5.1). In particular the Calderon projector for M cut

takes the form

PMcut =

(PM+ 00 PM−

). (5.5)

Warning. 1. There are two different decompositions of L2(E|∂Mcut), one comingfrom the ±i eigenspace decomposition of γ (2.5), and the second from the decomposition

∂M cut = N∐N (5.1). This leads to two different matrix representations of P ∈ Gr(A).

These two decompositions are compatible since γ = γ ⊕ (−γ), and so in fact one canwrite

L2(E|∂Mcut) = (Ei ⊕ E−i)⊕ (E−i ⊕ Ei).

2. Although P∆ ∈ Gr(A), it is not in Gr∞(A). This fact causes technical difficulties.3. It follows from (5.2) that if one parameterizes the collar of M− as ∂M− × [0, ε)

then γ is replaced by −γ and A is replaced by −A. This in particular means thatthe natural symplectic structure on L2(E|∂M−) is induced by −γ. Sometimes it will becrucial to distinguish between the map Φγ and the map Φ−γ (cf. (2.7)). The relationbetween the two is

Φ−γ(P ) = −Φγ(I − P )∗, P ∈ Gr(A). (5.6)

Lemma 5.1. Let D be a Dirac operator on M cut and suppose that Pt, 0 ≤ t ≤ 1, isa continuous path in Gr(A) and Qt, 0 ≤ t ≤ 1, is a continuous path in Gr(−A). Let

Bt =

(Pt 00 Qt

)

be the corresponding path in Gr(A). Then

SF(DBt,M cut)t∈[0,1] = SF(DPt

,M+)t∈[0,1] + SF(DQt,M−)t∈[0,1].

Proof. This follows from the fact that

L2(E,M cut) = L2(E,M+)⊕ L2(E,M−), (5.7)

and D,Bt preserve this splitting. Hence, DBt= DM+

Pt⊕DM−

Qt. Note that the splitting

(5.7) induces the splitting (5.1) by restricting to the boundary.

Notice that P ∈ Gr(A) if and only if I − P ∈ Gr(−A). Therefore, a particularlysymmetric family of boundary conditions for D acting on M cut is given by the imageof (P, I − P ) under the map (5.4).

Corollary 4.3 to the Scott–Wojciechowski theorem implies the following lemma.

20 PAUL KIRK AND MATTHIAS LESCH

Lemma 5.2. Let D be a Dirac operator over M and let P0, P1 ∈ Gr∞(A). Choosea smooth path Pt ∈ Gr∞(A), 0 ≤ t ≤ 1, from P0 to P1 and put

Qt :=

(Pt 00 I − Pt

)∈ Gr∞(A).

Then

η(DQ1,M cut)− η(DQ0

,M cut) = SF(DPt,M+)t∈[0,1] + SF(DI−Pt

,M−)t∈[0,1].

In particular, the quantity SF(DPt,M+)t∈[0,1] + SF(DI−Pt

,M−)t∈[0,1] is independent ofthe choice of the path Pt.

Proof. We know from Proposition 2.4 that Gr∞(A) is path connected. This assuresthe existence of a path Pt. Furthermore, notice that

η(DQt,M cut) = η(DPt

,M+) + η(DI−Pt,M−) (5.8)

since D and Qt preserve the splitting of L2(E|Mcut) = L2(E|M+)⊕ L2(E|M−).Lemma 3.4 and Corollary 4.3 imply that

η(DP1,M+)− η(DP0

,M+)− SF(DPt,M+)t∈[0,1]

=1

2

∫ 1

0

d

dt(η(DPt

))dt =1

2πi

∫ 1

0

d

dtlog detF

(Φ(Pt)Φ(P0)

∗)dt,

(5.9)

and

η(DI−P1,M−)− η(DI−P0

,M−)− SF(DI−Pt,M−)t∈[0,1]

=1

2

∫ 1

0

d

dt(η(DI−Pt

))dt =1

2πi

∫ 1

0

d

dtlog detF

(Φ−γ(I − Pt)Φ−γ(I − P0)

∗)dt.

(5.10)

Note that in (5.9) Φ is taken with respect to γ and in (5.10) Φ is taken with respect to−γ (cf. 3. of the warning above). In view of (5.6) we find

detF(Φ−γ(I − Pt)Φ−γ(I − P0)∗) = detF(Φ(Pt)

∗Φ(P0)) = detF(Φ(Pt)Φ(P0)∗), (5.11)

and consequently,

d

dtlog detF

(Φ−γ(I − Pt)Φ−γ(I − P0)

∗)= −

d

dtlog detF

(Φ(Pt)Φ(P0)

∗). (5.12)

Adding (5.9) and (5.10) and using (5.8) gives the desired formula.

For any P ∈ Gr(A) a natural path connecting P∆ and

(P 00 I − P

)is given by (cf.

[7, Sec. 3])

P (θ, P ) :=

(cos2(θ)P + sin2(θ)(I − P ) − cos(θ) sin(θ)

− cos(θ) sin(θ) cos2(θ)(I − P ) + sin2(θ)P

). (5.13)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 21

A straightforward calculation shows that ξ =

(ξ+ξ−

)∈ kerP (θ, P ) if and only if

cos(θ)Pξ+ = sin(θ)Pξ−,

sin(θ)(I − P )ξ+ = cos(θ)(I − P )ξ−.(5.14)

Lemma 5.3. Let P ∈ Gr(A). If cos(θ) 6= 0 then the map P (θ, P ) lies in Gr(A).Furthermore

P (0, P ) =

(P 00 I − P

)and P (π

4, P ) = P∆.

Proof. Fix a Lagrangian subspace L ⊂ kerA and let P+ = P+(L). The only part

which is not straightforward is the claim that (P (θ, P ), P+), P+ :=

(P+ 00 I − P+

)is

a Fredholm pair. We use the following criterion (cf. [6, Remark 3.5]).

Two orthogonal projections Q,R in a Hilbert space form a Fredholm pair(invertible pair) if and only if the operator QRQ+ (I −Q)(I −R)(I −Q)is Fredholm (invertible).

One calculates

P+P (θ)P+ + (I − P+)(I − P (θ, P ))(I − P+)

=(cos2(θ)(P+PP+ + (I − P+)(I − P )(I − P+)) +

sin2(θ)(P+(I − P )P+ + (I − P+)P (I − P+)))⊗

(I 00 I

)

≥ cos2(θ)(P+PP+ + (I − P+)(I − P )(I − P+))⊗

(I 00 I

).

(5.15)

Hence if cos(θ) 6= 0 then the pair (P (θ, P ), P+) is Fredholm (invertible) if the pair(P, P+) is Fredholm (invertible).

We emphasize that even if P ∈ Gr∞(A) then P (θ, P ) 6∈ Gr∞(A) if sin(θ) 6= 0. Thesignificance of the family P (θ, P ) stems from the fact that DP∆

is naturally unitarilyequivalent to D acting on the closed manifold M .

We first note some consequences of the existence of the family P (θ, P ) which do notmake use of η–functions.

Proposition 5.4. Let D be a Dirac operator over M and let P0, P1 ∈ Gr(A).Choose a smooth path Pt ∈ Gr(A), 0 ≤ t ≤ 1, from P0 to P1 and put, as in Lemma 5.2,

Qt :=

(Pt 00 I − Pt

)∈ Gr∞(A). (5.16)

Then

SF(DP (θ,P1),Mcut)θ∈[0,π

4] − SF(DP (θ,P0),M

cut)θ∈[0,π4] + SF(DQt

,M cut)t∈[0,1] = 0.

22 PAUL KIRK AND MATTHIAS LESCH

Proof. Note again that in view of Proposition 2.4 the space Gr(A) is path con-

nected. Using Pt one obtains a map H : [0, π4]× [0, 1] → Gr(A):

H(θ, t) = P (θ, Pt). (5.17)

Since H(π4, t) is the constant map at P∆, one sees that the path

θ 7→ H(θ, 0) = P (θ, P0), 0 ≤ θ ≤π

4,

is homotopic to the composite of the paths

t 7→ H(0, t) = Qt, 0 ≤ t ≤ 1,

and

θ 7→ H(θ, 1) = P (θ, P1).

The claim now follows from the homotopy invariance and additivity of the spectralflow.

Proposition 5.5. For the Calderon projectors PM+ of M+ and PMcut of M cut, thespace kerP (θ, PM+) ∩ ker γ(PMcut) is canonically isomorphic to imPM+ ∩ imPM−. Inparticular, its dimension is independent of θ ∈ [0, π

4]. Moreover,

SF(DP (θ,PM+ ),M

cut)θ∈[0,π4] = 0. (5.18)

Furthermore, if P ∈ Gr(A) and Pt, 0 ≤ t ≤ 1, is a smooth path in Gr(A) from P to theCalderon projector PM+ then

SF(DP (θ,P ),Mcut)θ∈[0,π

4] = SF(DQt

,M cut)t∈[0,1], (5.19)

where Qt = Pt ⊕ (I − Pt) as in (5.16).

Proof. By Lemma 2.3 the space kerP (θ, PM+) ∩ ker γ(PMcut) is isomorphic tokerDP (θ,P

M+). Hence, if we can show that kerP (θ, PM+) ∩ ker γ(PMcut) is independentof θ, then SF(DP (θ,P

M+),Mcut)θ∈[0,π

4] = 0.

Consider ξ =

(ξ+ξ−

)∈ kerP (θ, PM+)∩ ker γ(PMcut). In view of (5.5) and (5.14) this

means

cos(θ)PM+ξ+ = sin(θ)PM+ξ−,

0 = sin(θ)(I − PM+)ξ+ = cos(θ)(I − PM+)ξ−.(5.20)

Since cos(θ) 6= 0 we infer ξ− ∈ imPM+ ∩ imPM−.Conversely, given ξ− ∈ imPM+ ∩ imPM− put ξ+ := tan(θ)ξ−. Then (5.20) implies

that

(ξ+ξ−

)∈ kerP (θ, PM+) ∩ ker γ(PMcut).

(5.19) is an immediate consequence of (5.18) and Proposition 5.4.

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 23

Corollary 5.6. Let M be a split manifold and let D(t), a ≤ t ≤ b, be a smoothpath of Dirac type operators such that in a collar of the separating hypersurface we haveD(t) = γ( d

dx+ A(t)). Let PM+(t) be the corresponding family of Calderon projectors.

Then

SF(D(t))t∈[a,b] = SF(DI−PM+(t)(t),M

−)t∈[a,b].

Proof. We note that it was proved in [26] that PM+(t) is smooth. Consider thetwo parameter family of operators on M cut

(DP (θ,PM+(t))(t),M

cut)0≤θ≤π4,a≤t≤b.

By Proposition 5.5 for fixed t the dimension of the kernel of DP (θ,PM+ (t))(t) is indepen-

dent of θ. By the homotopy invariance of the spectral flow this implies

SF(DP (0,PM+ (t))(t),M

cut)t∈[a,b] = SF(DP (π4,P

M+(t))(t),Mcut)t∈[a,b].

Since P (π4) = P∆ the right hand side equals SF(D(t))t∈[a,b]. The left hand side equals

SF(DPM+ (t)(t),M

+)t∈[a,b] + SF(DI−PM+(t),M

−)t∈[a,b]

and since DP+M

(t)(t) is invertible its spectral flow vanishes and we reach the desired

conclusion.

Remark 5.7. We emphasize that we did not use η–invariants to prove Proposition5.4, Proposition 5.5, and Corollary 5.6. The only ingredients of the proof are the familyP (θ, P ) and basic properties of the spectral flow.

We now return to the discussion of η–invariants. Since DP∆is naturally unitarily

equivalent to D acting on the closed manifold M , we have for any P ∈ Gr(A)

η(DP (π4,P ),M

cut) = η(DP∆,M cut) = η(D,M). (5.21)

On the other hand, DP (0,P ) is the direct sum of DP acting on M+ and DI−P acting onM−. Therefore,

η(DP (0,P ),Mcut) = η(DP ,M

+) + η(DI−P ,M−). (5.22)

Hence, by Lemma 3.4 we have

η(D,M) = η(DP ,M+) + η(DI−P ,M

−)

+1

2

∫ π4

0

d

dθη(DP (θ,P ),M

cut)dt+ SF(DP (θ,P ))θ∈[0,π4].

(5.23)

Thus, in order to obtain a splitting theorem for the η–invariant one needs to under-stand the last two terms on the right hand side of (5.23). If P is the Calderon projectorof M+ or M− then by Proposition 5.5 the spectral flow term vanishes.

Consider now the Atiyah–Patodi–Singer projection P+ = P+(L) of (2.3). Thefollowing theorem is the main result of the article [7] by J. Bruning and M. Lesch ([7,Theorem 3.9], see also (3.68) of loc.cit. with T+ = −T ∗

− determined by the choice ofLagrangian L ⊂ kerA).

24 PAUL KIRK AND MATTHIAS LESCH

Theorem 5.8. Let P+ = P+(L) be the Atiyah–Patodi–Singer projection and letP (θ, P+)θ∈[0,π

4] the deformation (5.13) to the continuous transmission projection. Then

d

dθ(η(DP (θ,P+),M

cut) = 0.

In view of (5.23) we conclude from Theorem 5.8 that

η(D,M)− η(DP+,M+)− η(DI−P+,M−) = SF(DP (θ,P+),Mcut)θ∈[0,π

4]. (5.24)

Since the right hand side of (5.24) is an integer this formula implies the modZ gluingformula for the η–invariant (see [7] for a discussion of the history of this result). Notethat (5.24) is slightly weaker than Theorem 5.8.

Our strategy to obtain a useful splitting theorem for the η–invariant can now beexplained. On the one hand (5.24) gives a complete splitting formula for the η–invariantwith respect to Atiyah–Patodi–Singer boundary conditions, but it contains the (in gen-eral) uncomputable term SF(DP (θ,P+))θ∈[0,π

4]. On the other hand if we were to replace

P+ by the Calderon projector PM+, Proposition 5.5 shows that the corresponding spec-tral flow term vanishes. Thus Theorem 5.8 (or at least (5.24)) needs to be extendedto a more general class of projections, including the Calderon projector. One possiblestrategy would be to generalize the arguments of [7] to more general projections. Thismight be manageable but technically tedious. Here, we will use a simpler approachwhich shows slightly less. Lemma 5.2 and Proposition 5.5 lead to a generalization of(5.24) to projections in Gr∞(A). This is less than a generalization of Theorem 5.8 sincethe variation of the η–invariant with respect to the path P (θ, P ) might be non-zero.

Theorem 5.9. Let D be a Dirac operator on M and let N ⊂ M split M into M+

and M−. We assume that in a collar neighborhood [−ε, ε] × N of N , D has the formD = γ( d

dx+ A) as in (2.1). Let P ∈ Gr∞(A) and let Pt be a smooth path in Gr∞(A)

from P to the Calderon projector PM+. As in (5.16) put Qt := Pt ⊕ (I − Pt). Then

η(D,M) = η(DP ,M+) + η(DI−P ,M

−) + SF(DP (θ,P ),Mcut)θ∈[0,π

4]

= η(DP ,M+) + η(DI−P ,M

−) + SF(DQt,M cut)t∈[0,1].

(5.25)

In particular, if PM+ is the Calderon projector for M+ then

η(D,M) = η(DPM+,M+) + η(DI−P

M+,M−).

Proof. Fix a Lagrangian subspace L ⊂ kerA and choose a smooth path Rt ∈

Gr∞(A) from P to the Atiyah–Patodi–Singer projection P+ = P+(L). Set Rt :=Rt ⊕ (I − Rt). By Proposition 5.4 we have

SF(DP (θ,P ),Mcut)θ∈[0,π

4] = SF(DP (θ,P+),M

cut)θ∈[0,π4] + SF(DRt

,M cut)t∈[0,1].

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 25

Using Lemma 5.2 and (5.24) we obtain

η(D,M)− η(DP ,M+)− η(DI−P ,M

−)

= η(D,M)− η(DP+,M+)− η(DI−P+,M−) + η(DR1,M cut)− η(DR0

,M cut)

= SF(DP (θ,P+),Mcut)θ∈[0,π

4] + SF(DRt

,M cut)t∈[0,1]

= SF(DP (θ,P ),Mcut)θ∈[0,π

4].

(5.26)

This proves the first line of (5.25). The second line of (5.25) and the last assertionfollow from Proposition 5.5.

Notice that by symmetry the same argument also shows that

η(D,M) = η(DI−PM−,M+) + η(DP

M−,M−).

Applying Theorem 4.4 allows us to extend Theorem 5.9 as follows.

Theorem 5.10. Let D be a Dirac operator on M and let N ⊂M split M into M+

and M−. Then for P ∈ Gr∞(A) and Q ∈ Gr∞(−A) we have, with Φ = Φγ,

η(D,M)− η(DP ,M+)− η(DQ,M

−)

= − 12πi

tr log(Φ(P )Φ(PM+)∗)− 12πi

tr log(Φ(PM−)Φ(Q)∗)

+ 12πi

tr log(Φ(I − PM−)Φ(PM+)∗).

In particular,

η(D,M) = η(DPM+,M+) + η(DP

M−,M−) + 1

2πitr log(Φ(I − PM−)Φ(PM+)∗). (5.27)

Proof. It was remarked after (2.4) that PM+ ∈ Gr∞(A) and PM− ∈ Gr∞(−A).Consequently, I − PM− ∈ Gr∞(A) and hence I − PM− − PM+ is trace class.

Theorem 5.9 implies that η(D,M)− η(DP ,M+)− η(DQ,M

−) is equal to

− (η(DP ,M+)− η(DP

M+,M+))− (η(DQ,M

−)− η(DPM−,M−))

− (η(DPM−,M−)− η(DI−P

M+,M−)).

(5.28)

Applying Theorem 4.4 to the three summands in (5.28) and taking (5.6) into accountgives the assertion.

The formula (5.27) expresses the η–invariant of D onM in terms of two η–invariantsintrinsic to the two piecesM+ andM− of the decomposition and an “interaction” term.

6. Maslov index and winding number

In this section we compile the necessary material about the Maslov index and thewinding number. One important comment is that in constructing the various invariants(winding number, Maslov index, triple index, spectral flow, and the branch of thelogarithm) conventions must be chosen to set signs and to handle degenerate cases.In particular, care must be taken to ensure that the different possible conventionsare chosen compatibly. Thus, although some of the material we present here is ageneralization of ideas which appear in the literature, the subtleties arising in organizingthe conventions compatibly and extending the constructions from the finite–dimensionalto the infinite–dimensional context require the careful exposition we present.

26 PAUL KIRK AND MATTHIAS LESCH

6.1. Winding number. Let H be a complex Hilbert space and denote by U (H)the group of unitary operators on H . Similarly to (2.8), (2.9) we introduce the followingsubspaces:

U∗(H) :=U ∈ U (H)

∣∣ −1 6∈ specU,

UFred(H) :=U ∈ U (H)

∣∣ −1 6∈ specess U,

UK (H) :=U ∈ U (H)

∣∣ U − I is compact,

Utr(H) :=U ∈ U (H)

∣∣ U − I is trace class.

(6.1)

The spaces U∗(H) and UFred(H) are not groups. It is well–known that the inclusionUtr(H) → UK (H) is a homotopy equivalence and that UK (H) is homotopy equiv-alent to the infinite unitary group U (∞) = lim

n→∞U (n). Therefore, one has by Bott

periodicity

π2k(UK (H)) = π2k(Utr(H)) = 0,

π2k+1(UK (H)) = π2k+1(Utr(H)) ≃ Z,k = 0, 1, 2, ... (6.2)

Furthermore, the isomorphism π1(Utr(H)) → Z is given by the winding number. I.e. iff : [0, 1] → Utr(H) is a closed C1–path then

wind(f) :=1

2πi

∫ 1

0

tr(f(t)−1f ′(t))dt. (6.3)

Lemma 6.1.

1. The inclusion UK (H) → UFred(H) is a weak homotopy equivalence,2. For any U ∈ UFred(H) there exists a smooth path f : [0, 1] → UFred(H) such thatf(0)−I is of finite rank, f(1) = U , and such that dimker(f(t)+I) is independentof t.

Proof. 1. Let Q(H) := B(H)/K (H) be the Calkin algebra. Then the quotientmap σ : B(H) → Q(H) sends UFred(H) onto

u ∈ Q(H)

∣∣ −1 6∈ spec u=: U∗Q(H).

Moreover, UK (H) acts freely (from left and right) on the fibers. Thus we obtain afibration UK (H) → UFred(H) → U∗Q(H). The claim now follows since U∗Q(H) iscontractible.

To see the latter we note that for any C∗–algebra A the set u ∈ A | u unitary,−1 6∈spec u is contractible. The contraction is given by Ht(u) := exp(t log u), 0 ≤ t ≤ 1.This is well–defined since −1 6∈ spec u.

2. Let H = ker(U + I) ⊕ H1 =: H0 ⊕ H1. Then U splits into U = −IH0⊕ U and

−1 6∈ spec U . Now put f(t) := −IH0⊕ exp(t log U).

In view of this lemma the winding number (6.3) extends to a group isomorphism

wind : π1(UFred(H)) → Z.

Next we define the winding number for not necessarily closed paths in UFred(H).Namely, as it was noted in the previous proof the space U∗(H) is contractible. Hencethe natural map π1(UFred(H)) → π1(UFred(H),U∗(H)) is a bijection and thus we ob-tain a winding number defined for curves f : ([0, 1], 0, 1) → (UFred(H),U∗(H)).

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 27

More concretely, if f is such a curve then one chooses f : [0, 1] → U∗(H) with

f(0) = f(1) and f(1) = f(0). Then f ∗ f is a closed curve in UFred(H) and one

puts wind(f) := wind(f ∗ f). Since U∗(H) is contractible it is clear that wind(f) is

well–defined independently of the choice of f .Finally, we choose a convention to define the winding number for a curve whose

endpoints do not lie in U∗(H): let f : [0, 1] → UFred(H) be a continuous curve. −1 isan isolated point in the spectrum of f(t) since −1 6∈ specess(f(t)). We may thereforechoose an ε > 0 such that for all ϕ ∈ [−ε, ε], ϕ 6= 0, we have −1 6∈ spec(f(j)eiϕ), j = 0, 1.Now define

wind(f) := wind(fe−iε). (6.4)

The winding number has the following properties:

1. Path Additivity : Let f1, f2 : [0, 1] → UFred(H) be continuous paths with f2(0) =f1(1). Then

wind(f1 ∗ f2) = wind(f1) + wind(f2).

2. Homotopy invariance: Let f1, f2 be continuous paths in UFred. Assume that thereis a homotopy H : [0, 1]× [0, 1] → UFred such that H(0, t) = f1(t), H(1, t) = f2(t)and such that dim ker(H(s, 0) + I), dimker(H(s, 1) + I) are independent of s.Then wind(f1) = wind(f2).

3. If f : [0, 1] → Utr(H) is a C1–curve then

wind(f) =1

2πi

(∫ 1

0

tr(f(t)−1f ′(t))dt− tr(log f(1)) + tr(log f(0))), (6.5)

where the logarithm is normalized as in (4.11).

We note in passing that the winding number may be interpreted as a spectral flowacross −1 [4], [28]. Namely, the winding number of a path f : [0, 1] → UFred(H) can becalculated as follows: choose a subdivision 0 = t0 < t1 < ... < tn = 1 and 0 < εj < π,j = 0, ..., n−1, such that −eiϕ 6∈ specess(f(t)) for t ∈ [tj , tj+1] and |ϕ| ≤ εj and moreover−e±iεj 6∈ spec f(t) for t ∈ [tj, tj+1]. Then put

wind(f(t))tj≤t≤tj+1:= #(spec(f(tj+1)) ∩

−eiϕ

∣∣ 0 < ϕ < εj)

−#(spec(f(tj)) ∩−eiϕ

∣∣ 0 < ϕ < εj,

(6.6)

where eigenvalues are counted with multiplicity. Finally,

wind(f) =

n−1∑

j=0

wind(f(t))tj≤t≤tj+1. (6.7)

Definition 6.2. Let U ∈ UK (H) and V ∈ UFred(H). Then the double indexτw(U, V ) ∈ Z is defined as follows: choose continuous paths f : [0, 1] → UK (H), g :[0, 1] → UFred(H) such that f(0) = g(0) = I and f(1) = U , g(1) = V . Then putτw(U, V ) := wind(f) + wind(g) − wind(fg). τw(U, V ) is defined accordingly if U ∈UFred(H), V ∈ UK (H) or U, V ∈ UFred(H), UV ∈ UK (H).

28 PAUL KIRK AND MATTHIAS LESCH

Proposition 6.3. The double index τw is well–defined. It has the following prop-erties:

1. (Homotopy invariance) If f : [0, 1] → UK (H), g : [0, 1] → UFred(H) are continu-ous paths then

τw(f(1), g(1))− τw(f(0), g(0)) = wind(f) + wind(g)− wind(fg).

In particular, if dimker(f(t) + I), dimker(g(t)+ I) and dim ker(f(t)g(t)+ I) areindependent of t then τw(f(1), g(1)) = τw(f(0), g(0)).

2. If U, V ∈ Utr(H) then

τw(U, V ) =1

2πi

(tr logUV − tr logU − tr log V

). (6.8)

3. For any U ∈ UFred(H) we have

τw(I, U) = τw(U, I) = 0 and τw(U, U−1) = − dimker(U + I).

Proof. First note that if U ∈ UK (H) and V ∈ UFred(H) then since U − I is

compact one has specess(UV ) = specess(V ), in particular UV ∈ UFred(H). If f : [0, 1] →

UK (H), g : [0, 1] → UFred(H) are different paths with f(0) = g(0) = I, f(1) = U, g(1) =

V then consider the closed paths f ∗f− and g∗g−, where f− denotes the path f traversed

in the opposite direction. Since the pointwise product of closed paths (f ∗ f−)(g ∗ g−)

is homotopic to (f ∗ f−) ∗ (g ∗ g−) we find

0 = wind(f ∗ f−) + wind(g ∗ g−)− wind((f ∗ f−)(g ∗ g−))

= −wind(f) + wind(f)− wind(g) + wind(g) + wind(f g)− wind(fg).(6.9)

This shows that τw is well–defined. The homotopy invariance is straightforward fromthe definition and the homotopy invariance of the winding number.

2. This assertion is a consequence of (6.5).3. That τw(I, U) = τw(U, I) = 0 follows immediately from the definition.For U ∈ Utr(H) the third identity follows from Assertion 2. (note the normalization

(4.11) of log). If U is arbitrary we apply Lemma 6.1 2. and choose a continuous pathf : [0, 1] → UFred(H) such that f(1) = U, f(0) ∈ Utr(H) and such that dim ker(f(t)+I)is independent of t. The claim now follows from the homotopy invariance 1.

A priori τw cannot be defined on UFred(H) × UFred(H) (which might be desirable)since for U, V ∈ UFred(H) in general UV 6∈ UFred(H). Even if one assumes UV ∈UFred(H) it is in general not possible to choose paths f, g as above such that f(t)g(t) ∈UFred(H) for all t.

Corollary 6.4. Let f : [0, 1] → UFred(H) be a continuous path. Then

wind(f) + wind(f−1) = dimker(f(0) + I)− dimker(f(1) + I).

Proof. We apply Proposition 6.3 1. with g = f−1 and obtain using Proposition6.3 3.

wind(f) + wind(f−1) = τw(f(1), f(1)−1)− τw(f(0), f(0)

−1)= − dimker(f(1) + I) + dim ker(f(0) + I).

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 29

6.2. Maslov Index. Let (H, 〈 , 〉, γ) be a Hermitian symplectic Hilbert space (cf.Def. 2.8). Thus γ : H → H is a unitary map satisfying γ2 = −1 and the eigenspacesE±i := ker(γ ∓ i) have the same Hilbert space dimension. As in Section 2 we denote by

L :=L ⊂ H

∣∣ L closed subspace, γL = L⊥

the set of Lagrangian subspaces. As usual L ∈ L will be identified with the orthogonalprojection PL onto L. The image of an orthogonal projection P in H is Lagrangian ifand only if γPγ∗ = I − P . Similarly as in Section 2 we put

Gr(H) :=P ∈ B(H)

∣∣ P = P ∗, P 2 = P, γPγ∗ = I − P,

Gr(2)Fred(H) :=

(P,Q)

∣∣ P,Q ∈ Gr(H), (P,Q) are a Fredholm pair,

Gr(2)∗ (H) :=(P,Q)

∣∣ P,Q ∈ Gr(H), (P,Q) is an invertible pair,

Gr(2)K(H) :=

(P,Q)

∣∣ P,Q ∈ Gr(H), P −Q is compact.

(6.10)

Notice that, in contrast to the definition of Gr(A), there is no Fredholm assumptionabout elements of Gr(H). The corresponding spaces of Lagrangians are

L(2)Fred :=

(L1, L2)

∣∣ L1, L2 ∈ L , (L1, L2) is Fredholm,

L(2)∗ :=

(L1, L2) ∈ L

(2)∣∣ (L1, L2) is invertible

.

(6.11)

Recall that a pair of Lagrangian spaces (L1, L2) is Fredholm if L1 ∩ L2 is finite–dimensional and if L1 +L2 is closed with finite codimension, and that the pair (L1, L2)is invertible if L1 ∩ L2 = 0 and L1 + L2 = H .

We emphasize the confusing fact that (L1, L2) is Fredholm (resp. invertible) if andonly if the pair of projections (I − PL1

, PL2) is Fredholm (resp. invertible). There-

fore, a Fredholm pair of projections (P,Q) will sometimes be identified with the pair(kerP, imQ) of Lagrangian subspaces.

As in Lemma 2.6 one sees that with respect to the decomposition H = Ei⊕E−i eachP ∈ Gr(H) takes the form

P =1

2

(I Φ(P )∗

Φ(P ) I

), (6.12)

where Φ(P ) ∈ U (Ei, E−i). Moreover, the map

Φ : Gr(H) −→ U (Ei, E−i) (6.13)

is a diffeomorphism. Furthermore, the pair (P,Q) is Fredholm if and only ifΦ(P )Φ(Q)∗ ∈ UFred(E−i) and (P,Q) is invertible if and only if Φ(P )Φ(Q)∗ ∈ U∗(E−i)(cf. Lemma 2.6). Finally, P − Q is compact (resp. trace class) if and only ifΦ(P )Φ(Q)∗ ∈ UK (E−i) (resp. Utr(E−i)).

30 PAUL KIRK AND MATTHIAS LESCH

Proposition 6.5. There are diffeomorphisms

Gr(2)Fred(H) ∼= UFred(E−i)× U (Ei, E−i),

Gr(2)K(H) ∼= UK (E−i)× U (Ei, E−i),

Gr(2)∗ (H) ∼= U∗(E−i)× U (Ei, E−i),

Gr(2)K(H) ∩Gr(2)∗ (H) ∼=

(UK (E−i) ∩ U∗(E−i)

)× U (Ei, E−i).

In particular the identifications induce homotopy equivalences

(Gr(2)Fred(H),Gr(2)∗ (H)) ≃ (UFred(E−i),U∗(E−i))

(Gr(2)K(H),Gr(2)∗ (H) ∩GrK (H)) ≃ (UK (E−i),U∗(E−i) ∩ UK (E−i)).

Proof. In all four cases the diffeomorphism is given by

(P,Q) 7→ (Φ(P )Φ(Q)∗,Φ(P )).

By Kuiper’s Theorem [22] the space U (Ei, E−i) is contractible and hence we obtainthe claimed homotopy equivalences.

The Maslov Index [10], [26] is an integer invariant of Fredholm pairs of paths ofLagrangian subspaces. We discuss it in terms of the projection picture of Lagrangian

subspaces. Let (f, g) : [0, 1] → Gr(2)Fred(H) be a continuous path (i.e. a pair of paths in

Gr(H) such that (f(t), g(t)) is Fredholm for all t). The Maslov index Mas(f, g) is thealgebraic count of how many times ker f(t) passes through im g(t) along the path. Weuse the notation Mas(f, g),Mas(ker f, im g),Mas(im f, ker g) interchangeably. IndeedMas(ker f, im g) = Mas(γ ker f, γ im g) = Mas(im f, ker g).

The Maslov index has the following properties (cf. [26], [10]):

1. Path Additivity: Let (fj , gj) : [0, 1] → Gr(2)Fred(H), j = 1, 2, be continuous paths

with f2(0) = f1(1), g2(0) = g1(1) then

Mas((f1, g1) ∗ (f2, g2)) = Mas(f1, g1) + Mas(f2, g2).

2. Homotopy Invariance: Let (fj , gj) : [0, 1] → Gr(2)Fred(H), j = 0, 1, such that (f0, g0)

is homotopic (f1, g1) rel endpoints then

Mas(f0, g0) = Mas(f1, g1).

More generally, suppose that (F,G) is a homotopy from (f0, g0) =(F (−, 0), G(−, 0)) to (f1, g1) = (F (−, 1), G(−, 1)) and suppose thatdim(kerF (0, s)∩ imG(0, s)) and dim(kerF (1, s)∩ imG(1, s)) are independent ofs ∈ [0, 1]. Then Mas(f0, g0) = Mas(f1, g1).

3. Normalization: This is done in two steps. First one requires that on pathswith endpoints in Gr(2)∗ (H) the Maslov index induces a group isomorphism

π1(Gr(2)Fred(H),Gr(2)∗ (H)) → Z (since Gr(2)∗ (H) ≃ U∗(E−i) × U (Ei, E−i) is con-

tractible π1(Gr(2)Fred(H),Gr(2)∗ (H)) is indeed a group). This determines Mas on

paths with endpoints in Gr(2)∗ (H) up to a sign. The sign is chosen as follows: if

(P,Q) ∈ Gr(2)(H) then Mas(etγPe−tγ , Q)−ε≤t≤ε = dim(kerP ∩ imQ) for ε smallenough.

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 31

Secondly, if (f, g) : [0, 1] → Gr(2)Fred(H) is an arbitrary continuous path then

one chooses ε small enough such that the pairs (esγf(j)e−sγ, g(j)) are invertiblefor j = 0, 1, 0 < s ≤ ε. Then one puts

Mas(f, g) := Mas(eεγfe−εγ, g). (6.14)

Actually, the normalization property 3. determines the Maslov index completelyand it may be viewed as its definition. Properties 1. and 2. follow from 3. There existother conventions for dealing with paths whose endpoints do not lie in Gr(2)∗ (H) andnot all of these conventions satisfy path additivity.

The discussion of the Maslov index works as well when the Hermitian symplecticHilbert space H is finite-dimensional. In this case the Fredholm condition is vacuousand the Maslov index is defined on Gr(2)(H) = Gr(H)×Gr(H). We will use the Maslovindex in both contexts; in the infinite–dimensional setting with H = L2(E|∂X) and inthe finite–dimensional context with H = kerA.

Theorem 6.6. For a continuous path (f, g) in Gr(2)Fred(H) the Maslov index is related

to the winding number by the equation

Mas(f, g) = −wind(Φ(f)Φ(g)∗). (6.15)

Proof. In view of Proposition 6.5 the right hand side of (6.15) induces a group

isomorphism π1(Gr(2)Fred(H),Gr(2)∗ (H)) → Z. It remains therefore to check the sign

convention and the convention for paths with endpoints not in Gr(2)∗ (H). Let (P,Q) ∈

Gr(2)Fred(H). Then, by definition, Mas(etγPe−tγ , Q)−ε≤t≤ε = dim(kerP ∩imQ) for ε small

enough. A straightforward calculation shows

Φ(esγPe−sγ) = e−2sγΦ(P ) (6.16)

and thus for ε > 0 small enough we have, in view of Lemma 2.6 (3),

wind(Φ(esγPe−sγ)Φ(Q)∗)−ε≤s≤ε = dim(kerP ∩ imQ) wind(−e−2is)−ε≤s≤ε

= − dim(kerP ∩ imQ).(6.17)

To check the convention for paths with endpoints not in Gr(2)∗ (H) we consider the paths(esγPe−sγ, Q),−ε ≤ s ≤ 0, resp. 0 ≤ s ≤ ε. By definition we have for δ > 0 smallenough

Mas(esγPe−sγ, Q)−ε≤s≤0 = Mas(e(δ+s)γPe−(δ+s)γ , Q)−ε≤s≤0

= Mas(etγPe−tγ, Q)−ε+δ≤s≤+δ = dim(kerP ∩ imQ),(6.18)

and, analogously,

Mas(esγPe−sγ, Q)0≤s≤ε = 0. (6.19)

32 PAUL KIRK AND MATTHIAS LESCH

According to our convention for the winding number we have, on the other hand,

wind(−e−2is)−ε≤s≤0 = −1,

wind(−e2is)0≤s≤ε = 0.(6.20)

In view of (6.17) the proof is complete.

Corollary 6.7. Let (f, g) be a continuous path in Gr(2)Fred(H).

(1) The Maslov index Mas−γ with respect to the opposite symplectic structure is re-lated to Masγ as follows: Mas−γ(f, g) = Masγ(g, f).

(2) Masγ(f, g) + Masγ(g, f) = dim(ker f(1) ∩ im g(1))− dim(ker f(0) ∩ im g(0)).

Proof. (1) In view of (5.6) and the previous theorem we find Mas−γ(f, g) =−wind(Φ−γ(f)Φ−γ(g)

∗) = −wind(Φγ(f)∗Φγ(g)) = wind(Φγ(g)Φγ(f)

∗) = Masγ(g, f).(2) Using the previous Theorem and Corollary 6.4 we obtain (we write Mas instead

of Masγ)

Mas(f, g) + Mas(g, f) = −wind(Φ(f)Φ(g)∗)− wind((Φ(f)Φ(g)∗)−1)= dim(ker f(1) ∩ im g(1))− dim(ker f(0) ∩ im g(0)).

Finally we construct a version of the Maslov triple index in our context. TheMaslov triple index as defined in (cf. [10, Sec. 8]) cannot be generalized to the presentinfinite–dimensional setting. The reason is simply that interesting triples of Lagrangiansubspaces L1, L2, L3 such that (L1, L2), (L2, L3), (L3, L1) are all Fredholm pairs are hardto find.

However, motivated by [10, Prop. 8.4] we can construct a variant τµ of the Maslovtriple index which is related to the double index τw: consider continuous paths f, g, h :

[0, 1] → Gr(H) such that (f, g), (g, h), (f, h) map into Gr(2)Fred(H) and such that f − g or

g − h or f − h maps into the set of compact operators. If, say, f(t) − g(t) is compactfor all t then, of course, it suffices to assume that (f(t), h(t)) is Fredholm for all t. TheFredholmness of (f(t), g(t)), (g(t), h(t)) then follows. Now in view of Theorem 6.6 andProposition 6.3 we find

Mas(f, g) + Mas(g, h)−Mas(f, h)

= −wind(Φ(f)Φ(g)∗)− wind(Φ(g)Φ(h)∗) + wind(Φ(f)Φ(h)∗)

= −τw(Φ(f(1))Φ(g(1))∗,Φ(g(1))Φ(h(1))∗)

+ τw(Φ(f(0))Φ(g(0))∗,Φ(g(0))Φ(h(1))∗).

(6.21)

This motivates the following definition.

Definition 6.8. Let P,Q,R ∈ Gr(H) such that (P,Q), (Q,R), (P,R) are Fredholmand at least one of the differences P −Q,Q− R,P − R is compact. Then we set

τµ(P,Q,R) := −τw(Φ(P )Φ(Q)∗,Φ(Q)Φ(R)∗). (6.22)

and call τµ the triple index of (P,Q,R).

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 33

The triple index τµ inherits properties from τw in a straightforward way. For exam-ple, one has the following.

Lemma 6.9. Let P,Q,R ∈ Gr(H) such that P −Q,Q− R are trace class. Then

τµ(P,Q,R) =1

2πi

(tr log(Φ(P )Φ(Q)∗) + tr log(Φ(Q)Φ(R)∗)

− tr log(Φ(P )Φ(R)∗)).

(6.23)

We will have occasion below to use the homotopy invariance of the triple index.

Lemma 6.10. Let P,Q,R : [0, 1] → Gr(H) be paths in Gr(H) so that

(P,Q), (Q,R), (P,R) map into Gr(2)Fred(H) and at least one of the differences

P − Q,Q − R,P − R maps into the set of compact operators. Suppose furtherthat dim(kerP (t) ∩ imQ(t)), dim(kerQ(t) ∩ imR(t)), and dim(kerP (t) ∩ imR(t)) areindependent of t. Then

τµ(P (0), Q(0), R(0)) = τµ(P (1), Q(1), R(1)).

Proof. By (6.21) we have

τµ(P (0), Q(0), R(0))− τµ(P (1), Q(1), R(1))

= Mas(P,Q) + Mas(Q,R)−Mas(P,R).(6.24)

Now the claim follows immediately from the homotopy invariance of the Maslov index.

We defined the triple index in such a way that formulas become short. A drawbackof this is that τµ is not antisymmetric in the variables, as the following propositionshows.

Proposition 6.11. Let P,Q,R ∈ Gr(H) such that (P,Q), (Q,R), (P,R) are Fred-holm and at least one of the differences is compact. Then

τµ(P,R,Q) =− τµ(P,Q,R) + dim(kerQ ∩ imR),

τµ(Q,P,R) =− τµ(P,Q,R) + dim(kerP ∩ imQ),

τµ(R,Q, P ) =− τµ(P,Q,R) + dim(kerP ∩ imQ)

+ dim(kerQ ∩ imR)− dim(kerP ∩ imR).

(6.25)

Moreover,

τµ(P, P,Q) = τµ(Q,P, P ) = 0 and τµ(P,Q, P ) = dim(kerP ∩ imQ). (6.26)

Proof. We prove (6.26) first. From Proposition 6.3 and the definition of τµ weinfer

τµ(P, P,Q) = −τw(I,Φ(P )Φ(Q)∗) = 0,

τµ(Q,P, P ) = −τw(Φ(Q)Φ(P )∗, I) = 0,

τµ(P,Q, P ) = −τw(Φ(P )Φ(Q)∗, (Φ(P )Φ(Q)∗)−1) = dim(kerP ∩ imQ).

(6.27)

34 PAUL KIRK AND MATTHIAS LESCH

To prove (6.25) we assume, without loss of generality, that Q − R is compact. Letf(t) := (1− t)Q + tR, 0 ≤ t ≤ 1. Then we obtain from Corollary 6.7 and (6.26)

τµ(P,R,Q) = τµ(P,R,Q)− τµ(P,Q,Q) = Mas(P, f) + Mas(f,Q)

= Mas(P, f)−Mas(Q, f) + dim(kerQ ∩ imR)

= −τµ(P,Q,R) + dim(kerQ ∩ imR),

τµ(Q,P,R) = Mas(P, f)−Mas(Q, f) + τµ(Q,P,Q)

= −τµ(P,Q,R) + dim(kerP ∩ imQ),

τµ(R,Q, P ) = Mas(f,Q)−Mas(f, P )

= −Mas(Q, f) + Mas(P, f) + dim(kerP ∩ imQ)

+ dim(kerQ ∩ imR)− dim(kerP ∩ imR).

(6.28)

6.3. Symplectic reduction. We conclude this section with a discussion of sym-plectic reduction in our infinite–dimensional context. We will use symplectic reductionin Section 8.

Let (H, 〈., .〉, γ) be a Hermitian symplectic Hilbert space with symplectic formω(x, y) = 〈x, γy〉. For a subspace U ⊂ H the annihilator of U is defined to be

Ann(U) :=x ∈ H

∣∣ ∀y ∈ U ω(x, y) = 0= (γU)⊥.

A subspace U ⊂ H is called isotropic if U ⊂ Ann(U).Assume for the moment that H is finite–dimensional and that Ann(U) ⊂ U . Then ω

induces a symplectic structure on the quotient U/Ann(U) in a natural way. Moreover,if L ⊂ H is Lagrangian then RU(L) := L∩U/L∩Ann(U) is Lagrangian in U/Ann(U).RU(L) is called the symplectic reduction of L.

Proposition 6.12. Let (H, 〈., .〉, γ) be a Hermitian symplectic Hilbert space, U ⊂H a closed subspace with Ann(U) ⊂ U .

Suppose that L ⊂ H is a Lagrangian subspace such that L + Ann(U) is a closedsubspace of H. Then (U ∩ γU, 〈., .〉, γ) is a Hermitian symplectic Hilbert space and theorthogonal projection

PL,U := projU∩γU : L ∩ U −→ U ∩ γU

has closed range isomorphic to L ∩ U/L ∩ Ann(U). Moreover, RU(L) = imPL,U isLagrangian in U ∩ γU .

RU (L) is called the symplectic reduction of L with respect to U .

Remark 6.13.

1. For (U ∩ γU, 〈., .〉, γ) to be Hermitian symplectic it is crucial that there is atleast one Lagrangian subspace L ⊂ H with L+Ann(U) closed. To illustrate theproblem start with an infinite–dimensional Hermitian symplectic Hilbert space

(H, 〈., .〉, γ). Let H := H ⊕H1, where H1 is another Hilbert space, and put γ :=

γ⊕i. Furthermore, pick a symplectic subspace L ⊂ H and put U := L⊕H1 ⊂ H .

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 35

Then Ann(U) = L⊕ 0 and U ∩ γU = 0⊕H1. Since γ acts by multiplication by ion U ∩ γU we conclude that (U ∩ γU, 〈., .〉, γ) is not Hermitian symplectic. From

the proposition we infer that for each Lagrangian subspace K ⊂ H the spaceK +Ann(U) is not closed.

2. Proposition 6.12 in particular applies if (L,Ann(U)) is a Fredholm pair of sub-spaces.

3. The assignment L 7→ RU(L) is not continuous, but is continuous along paths Lt

so that dim(Lt ∩ Ann(U)) is constant. These facts are well–known and we omitthe examples.

Proof. Certainly U ∩ γU is a Hilbert space and γ leaves U ∩ γU invariant. Ifwe can prove that imPL,U is Lagrangian in U ∩ γU then from Lemma 2.7 we inferdim(ker(γ + i) ∩ U ∩ γU) = dim(ker(γ − i) ∩ U ∩ γU).

What remains, therefore, is to prove the second part of Proposition 6.12 withoutusing the fact that dim(ker(γ + i) ∩ U ∩ γU) = dim(ker(γ − i) ∩ U ∩ γU).

We note first that we have an orthogonal direct sum decomposition

Ann(U)⊕ (U ∩ γU) = U. (6.29)

Also, imPL,U is an isotropic subspace of U ∩ γU . In fact, if x ∈ L ∩ U then,since L is Lagrangian, 〈x, γx〉 = 0. Writing x = ξ + η, ξ ∈ U ∩ γU, η ∈ Ann(U) then0 = 〈x, γx〉 = 〈ξ, γx〉 = 〈ξ, ξ〉 = 〈PL,Ux, γPL,Ux〉.

Next consider ξ ∈ U ∩ γU such that γ(ξ) ⊥ imPL,U . Thus for all x ∈ L ∩ U we

have 〈γ(ξ), x〉 = 〈γ(ξ), PL,Ux〉 = 0. Hence γ(ξ) ∈ (L ∩ U)⊥ = L⊥ + U⊥ = γ(L) + U⊥

and consequently, since L + Ann(U) is closed, ξ ∈ L + Ann(U). We may write ξ =l + η, l ∈ L, η ∈ Ann(U). From ξ ∈ U, η ∈ Ann(U) ⊂ U we infer l ∈ L ∩ U and henceξ = PL,U(l) ∈ imP (L).

Summing up we have proved γ((imPL,U)⊥) ⊂ imPL,U . Since imPL,U is isotropic this

implies imPL,U = γ((imPL,U)⊥). Thus imPL,U is a Lagrangian (in particular closed)

subspace of U ∩ γU .From (6.29) it is now clear that imPL,U is isomorphic to L ∩ U/L ∩ Ann(U).

7. Splittings of manifolds and the η–invariant II

For the proof of Theorem 5.9, Lemma 5.2 was crucial. The proof of Lemma 5.2depends on the Scott–Wojciechowski theorem 4.1. In this section we want to give proofsof Lemma 5.2 and Theorem 5.9 which are independent of the Scott–Wojciechowskitheorem and which apply to all P ∈ Gr(A). We only use (a mild generalization of)Theorem 5.8. Moreover, we derive generalizations of two results due to L. Nicolaescu[26]. This in turn leads to a nicer version of the splitting formula for the η–invariantwhich involves our version of the Maslov triple index.

We first introduce a setting which slightly generalizes the one described in Section2. Let X be a compact Riemannian manifold with boundary ∂X = Y

∐Z, i.e. the

boundary is a disjoint union of two (not necessarily connected) manifolds. We assumethat in collars U = UY and UZ we have D = γY (

ddx

+AY ) (resp. D = γZ(ddx

+AZ)) andthat the ±i–eigenspaces of γY (γZ) acting on kerAY (kerAZ) have the same dimension.

36 PAUL KIRK AND MATTHIAS LESCH

The latter does not follow as in section 2; rather it is an assumption. We fix onceand for all a PZ ∈ Gr(AZ). Then we can define the Calderon projector (relative toPZ) in Gr(AY ). Write again A instead of AY . It will be convenient to address Y, Z asboundary components although Y, Z are not assumed to be connected.

The results of Sections 2 to 5 generalize verbatim to this more general setting.Also Theorem 5.8 applies to this setting since all proofs work locally in a collar of theseparating hypersurface. The advantage of this setting is that it allows in particular toglue cylinders of the form [0, ε]×N with different boundary conditions on the ends toa manifold.

Lemma 7.1 (cf. [24, Lemma 2.5]). LetM = [0, ε]×N and D = γ( ddx+A) as before.

Moreover, let P,Q ∈ Gr(A) and denote by DP,Q be the operator obtained by imposingthe boundary condition P at 0 × N and I − Q at ε × N . Then λ ∈ specDP,Q ifand only if −λ ∈ specDQ,P . In particular,

η(DP,Q) = −η(DQ,P ), dimkerDP,Q = dimkerDQ,P .

Proof. The proof is exactly the same as the proof of [24, Lemma 2.5]. Namely,the isometry

T : L2([0, ε], L2(E|N)) −→ L2([0, ε], L2(E|N)), T f(x) := γf(ε− x)

maps the domain of DP,Q onto the domain of DQ,P and it anticommutes with D. HenceT ∗DP,QT = −DQ,P and we are done.

Now let M be a Riemannian manifold with boundary containing a separating hy-persurface N ⊂ (M \ ∂M). Let D be a Dirac operator as in Section 5; i.e. in a collarneighborhood [−ε, ε]×N of N , D has the form D = γ( d

dx+A) as in (2.1). Moreover, we

assume that the ±i–eigenspaces of γ acting on kerA have the same dimension. DefineM cut,M± as in Section 5. We assume that on the boundary components of (∂M±) \Nself–adjoint boundary projections have been fixed once and for all.

Lemma 7.2. For any P ∈ Gr(A) we have η(D,M)−η(DP ,M+)−η(DI−P ,M

−) ∈ Z.

Proof. Denote byM cutε the manifold with boundary obtained by removing [−ε, ε]×

N from M . As in Lemma 7.1 for P,Q ∈ Gr(A) we denote by η(DP,Q, [−ε, ε]×N) theη–invariant of the operator on [−ε, ε]×N obtained from D by imposing the boundarycondition P at −ε × N and the boundary condition I − Q at ε × N . The modZgluing formula for the η–invariant (5.24) then implies

η(D,M) ≡ η(DP+⊕(I−P+),Mcutε ) + η(DP+,P+, [−ε, ε]×N)modZ (7.1)

for P+ = P+(L) the Atiyah–Patodi–Singer projection with respect to a Lagrangiansubspace L ⊂ kerA. One easily checks that kerDP+,P+ = 0, hence Lemma 7.1implies

η(DP+,P+, [−ε, ε]×N) = 0. (7.2)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 37

Also by Lemma 7.1

η(DP+,P , [−ε, 0]×N) + η(DP,P+, [0, ε]×N)

=1

2dim ker(DP+,P , [−ε, 0]×N) +

1

2dim ker(DP,P+, [0, ε]×N)

= dimker(DP,P+, [0, ε]×N) ∈ Z.

(7.3)

Plugging this into (7.1) and applying again the modZ splitting formula for the η–invariant we get

η(D,M) ≡ η(DP+⊕(I−P+),Mcutε ) + η(DP+,P , [−ε, 0]×N) + η(DP,P+, [0, ε]×N)

≡ η(DP ,M+) + η(DI−P ,M

−)modZ.(7.4)

Lemma 7.3. Lemma 5.2 holds for all P0, P1 ∈ Gr(A).

Proof. We freely use the notations of Lemma 5.2 and its proof. By Lemma 7.2 wehave for all t

η(DQt,M cut)− η(DQ0

,M cut)

= (η(DQt,M cut)− η(D,M))− (η(DQ0

,M cut)− η(D,M)) ∈ Z.(7.5)

Henced

dtη(DQt

,M cut) = 0 (7.6)

and the assertion follows from Lemma 3.4 and Lemma 5.1.

Now we can prove the following considerable generalization of the splitting formulafor the η–invariant. In Theorem 5.9 we assumed that P ∈ Gr∞(A). In the followingtheorem we only require P ∈ Gr(A).

Theorem 7.4. The statement of Theorem 5.9 remains valid if P ∈ Gr(A) and Pt

is a smooth path in Gr(A) from P to the Calderon projector.

Proof. The proof is exactly the same as the one of Theorem 5.9. One only has toinvoke Lemma 7.3 instead of Lemma 5.2.

We next present generalizations of two results due to L. Nicolaescu [26].

Theorem 7.5. Let X be a manifold with boundary and D(t), a ≤ t ≤ b, a smoothfamily of Dirac operators. We assume that in a collar of the boundary D takes the formγ( d

dx+ A(t)) as before. Let P (t) ∈ Gr(A(t)) be a smooth family. Denote by PX(t) the

Calderon projectors of D(t), and LX(t) = imPX(t) the Cauchy data spaces. Then

SF(DP (t)(t))t∈[a,b] = Mas(P (t), PX(t))t∈[a,b] = Mas(kerP (t), LX(t))t∈[a,b].

Note that γ is assumed to be constant. This is essential. Note that in [14, Theorem4.3] the collar of ∂X was parametrized as (−ε, 0]× ∂X . Their formula is obtained byinvoking Corollary 6.7.

38 PAUL KIRK AND MATTHIAS LESCH

Proof. We first consider the case P (t) ∈ Gr∞(A(t)). Since DPX(t)(t) is invertible,its spectral flow vanishes. We apply Lemma 3.4, Theorem 4.4, (6.5), and Theorem 6.6to calculate

SF(DP (t)(t))t∈[a,b] = SF(DP (t)(t))t∈[a,b] − SF(DPX(t)(t))t∈[a,b]= η(DP (b)(b))− η(DPX(b)(b))− η(DP (a)(a)) + η(DPX(a)(a))

∫ b

a

d

dt

(η(DP (t)(t))− η(DPX(t)(t))

)dt

= 12πi

tr log(Φ(P (b))Φ(PX(b)∗))− 1

2πitr log(Φ(P (a))Φ(PX(a)

∗))

∫ b

a

12πi

d

dttr log(Φ(P (t))Φ(PX(t))

∗)dt

= −wind(Φ(P (t))Φ(PX(t))∗)t∈[a,b] = Mas(P (t), PX(t))t∈[a,b].

Now suppose that P (t) is arbitrary. Choose smooth paths P0(t) in Gr(A(0)) andP1(t) ∈ Gr(A(1)) such that P0(0) ∈ Gr∞(A(0)), P0(1) = P (0), P1(0) = P (1), P1(1) ∈Gr∞(A(1)) and such that

dim(kerP0(t) ∩ imPX(0)) and dim(kerP1(t) ∩ imPX(1)) (7.7)

are independent of t. The existence of P0, P1 follows from Lemma 6.1 by consideringΦ(P (j))Φ(PX(j))

∗, j = 0, 1. In view of (7.7) and Lemma 2.3 the dimension of the ker-nels of DP0(t)(0) and DP1(t)(1) are constant and hence the spectral flow of DP0(t)(0) andDP1(t)(1) vanishes. We may therefore compose the pathsDP0(t)(0), DP (t)(t), DP1

(t) with-out changing the spectral flow. Also Mas(P0(t), PX(0)) = Mas(P1(t), PX(1)) = 0 in viewof (7.7). In sum, without loss of generality we may assume that the family P (t) satis-fies P (0) ∈ Gr∞(A(0)), P (1) ∈ Gr∞(A(1)). Now consider the path Φ(P (t))Φ(PX(t))

∗ inUFred. In view of Lemma 6.1 this path is homotopic rel endpoints to a path f(t) ∈ U∞.

Putting P (t) := Φ−1(f(t)Φ(PX(t))) ∈ Gr∞(A(t)) we see that (P (t), PX(t)) is homotopic

rel endpoints to the path (P (t), PX(t)). Since homotopies with fixed endpoints neitherchange the spectral flow nor the Maslov index we find

SF(DP (t)(t))t∈[a,b] = SF(DP (t)(t))t∈[a,b] = Mas(P (t), PX(t))t∈[a,b]= Mas(P (t), PX(t))t∈[a,b].

We also give a generalization of Nicolaescu’s theorem for closed manifolds. Theresult in the following form was first proven in [13].

Theorem 7.6. Let M be a split manifold as in Section 5 and let D(t), a ≤ t ≤ b, bea smooth path of Dirac type operators such that in a collar of the separating hypersurfacewe have D(t) = γ( d

dx+ A(t)). Then

SF(D(t))t∈[a,b] = Masγ(PM−(t), I − PM+(t))t∈[a,b] = Mas(LM−(t), LM+(t))t∈[a,b].

Proof. Corollary 5.6 states that

SF(D(t))t∈[a,b] = SF(DI−PM+(t)(t),M

−)t∈[a,b]. (7.8)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 39

Applying Theorem 7.5 to the right hand side of (7.8) and using Corollary 6.7 yields

SF(DI−PM+ (t)(t),M

−)t∈[a,b] = Mas−γ(I − PM+(t), PM−(t))t∈[a,b]= Masγ(PM−(t), I − PM+(t))t∈[a,b],

finishing the proof.

Notice that the proof of Theorem 7.6 does not rely on Theorem 5.9, and in particulardoes not use the result of [7].

Finally, we state the following nicer version of the gluing formula for the η–invariant.We emphasize that the term τµ(I−PM−

, P, PM+), which was defined in Subsection 6.2,is an integer invariant which is defined completely in terms of the Hermitian symplecticstructure on L2(E|N).

Theorem 7.7. In the situation of Theorem 5.9, let P ∈ Gr(A). Then

η(D,M) = η(DP ,M+) + η(DI−P ,M

−)− τµ(I − PM−, P, PM+).

Proof. We note again that I−PM− −PM+ is trace class (cf. the proof of Theorem5.10). Thus I−PM− −PM+ is compact and hence the triple index τµ(I−PM−, P, PM+)is well–defined for any P ∈ Gr(A).

Let Pt, 0 ≤ t ≤ 1, be a smooth path in Gr(A) from P to the Calderon projectorPM+. Notice that Masγ(I−PM− , PM+) = 0 since I−PM− and PM+ are constant paths.From Theorem 7.4, Theorem 7.5, (6.21), Corollary 6.7, and Proposition 6.11 we infer

η(D,M)− η(DP ,M+)− η(DI−P ,M

−)

= SF(DPt,M+)t∈[0,1] + SF(DI−Pt

,M−)t∈[0,1]= Masγ(Pt, PM+)t∈[0,1] +Mas−γ(I − Pt, PM−)t∈[0,1]= Masγ(Pt, PM+)t∈[0,1] +Masγ(I − PM−, Pt)t∈[0,1] −Masγ(I − PM−, PM+)t∈[0,1]= τµ(I − PM−, PM+, PM+)− τµ(I − PM−

, P, PM+)= −τµ(I − PM−

, P, PM+).

8. Adiabatic stretching and applications to the Atiyah-Patodi-Singer

ρ-invariant

For the purpose of computation, one weakness of the splitting formulas of Theorems5.9, 5.10, and 7.7 is that it is difficult in practice to identify the Calderon projector.In many applications it is more convenient to work with the Atiyah–Patodi–Singerprojection P+(L), or at least some finite rank perturbation of P+(L), as a boundarycondition. According to Theorem 5.9, this requires knowing the spectral flow of DPt

and DI−Ptalong a path Pt starting at the Calderon projector and ending at P+(L).

A natural choice of such a path is the path obtained by stretching the collar neigh-borhood of the separating surface. According to a theorem of Nicolaescu [26], theCalderon projector limits to a projection of the form P>ν + projL, where P>ν is theprojection to the span of the eigenvectors of A with eigenvalues greater than ν and

40 PAUL KIRK AND MATTHIAS LESCH

L is a Lagrangian subspace of the (finite–dimensional) span of eigenvectors of A witheigenvalues in the range [−ν, ν]. The number ν is the non–resonance level [26] of Dacting onM+ and in particular is zero if and only if there are no L2 solutions to Dφ = 0on the manifold obtained fromM+ by adding an infinite collar. If ν = 0, then the limitof the Calderon projector is an Atiyah–Patodi–Singer projection P+(V ) for a particularLagrangian V ⊂ kerA.

This approach works particularly well to study the odd signature operator andthe Atiyah–Patodi–Singer ρα invariant [2], since the effect of the Riemannian metric isminimized in this important case. We present the details. The approach can be made towork for arbitrary Dirac operators as well, however additional correction terms appearcorresponding to the 1–parameter family of operators acting onM andM± as the collarof the separating hypersurface is stretched to infinity. We will make some commentsabout the case of general Dirac operators at the end of this section.

8.1. The odd signature operator. Let X be a compact manifold of dimension2n + 1, with (possibly empty) boundary ∂X2n. Assume a collar of ∂X is isometric to[0, ε) × ∂X . Let α : π1(X) → U(n) be a representation. To α one can assign a flatvector bundle, that is, a Cn bundle E → X together with a flat connection B on Eso that the holonomy representation of B is equal to α. If ∂X is non–empty, we mayassume, by gauge transforming B if necessary, that B is in temporal gauge on the collar,in other words there is a flat U(n) connection b on E|∂X so that the restriction of B tothe collar [0, ε)× ∂X is of the form

B[0,ε)×∂X = q∗(b),

where q : [0, ε)× ∂X → ∂X is the projection to the second factor.Let dB : Ωp(X ;E) → Ωp+1(X ;E) and db : Ω

p(∂X ;E|∂X) → Ωp+1(∂X ;E|∂X) denotethe associated coupled DeRham operators. Note that d2B and d2b are zero since B andb are flat. The cohomology of the complex (Ω∗(X ;E), dB) (resp. (Ω∗(∂X ;E|∂X), db))is identified via the DeRham theorem with the singular cohomology H∗(X ;Cn

α) (resp.H∗(∂X ;Cn

α)), where Cnα denotes the local coefficient system determined by the repre-

sentation α.The odd signature operator on X coupled to the flat connection B is the operator

DB : ⊕pΩ2p(X ;E) → ⊕

pΩ2p(X ;E)

defined by

DB(β) = in+1(−1)p−1(∗dB − dB∗)(β) for β ∈ Ω2p(X ;E),

where ∗ : Ωk(X ;E) → Ω2n+1−k(X ;E) denotes the Hodge ∗ operator (see [2]).The operator DB is a symmetric Dirac operator. Its square is the twisted Laplacian

acting on even bundle–valued forms:

D2B = d∗BdB + dBd

∗B.

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 41

In particular DB is self–adjoint if X has empty boundary and in that case its kernelcan be identified with the twisted cohomology group ⊕pH

2p(X ;Cnα) by the Hodge and

DeRham theorems. This implies that the dimension of the kernel of DB is independentof the choice of Riemannian metric if X is closed.

Define a restriction map

r : ⊕pΩ2p(X ;E) → ⊕

kΩk(∂X ;E|∂X)

by the formular(β) = i∗(β) + i∗(∗β)

where i : ∂X → X denotes the inclusion of the boundary.To avoid confusion we denote the Hodge ∗ operator on the boundary by ∗, thus

∗ : Ωk(∂X ;E|∂X) → Ω2n−k(∂X ;E|∂X).

We use ∗ to defineγ : ⊕

kΩk(∂X ;E|∂X) → ⊕

kΩk(∂X ;E|∂X)

by

γ(β) =

in+1(−1)p−1∗ β if β ∈ Ω2p(∂X ;E|∂X),

in+1(−1)n−q∗ β if β ∈ Ω2q+1(∂X ;E|∂X).

Finally, we define the operator

Ab : ⊕kΩk(∂X ;E|∂X) → ⊕

kΩk(∂X ;E|∂X)

by

Ab(β) =

−(db∗+ ∗db)β if β ∈ ⊕k Ω

2k(∂X ;E|∂X),

(db∗+ ∗db)β if β ∈ ⊕k Ω2k+1(∂X ;E|∂X).

The following facts are routine to verify.

1. Ab is a self–adjoint Dirac operator on ∂X .2. r induces an identification Φ : ⊕pΩ

2p([0, ε)×∂X ;E) → C∞([0, ε),⊕k Ωk(∂X ;E))

which is isometric with respect to the L2–structures. Moreover,

ΦDBΦ∗ = γ( ∂

∂x+ Ab), (8.1)

where x denotes the collar coordinate.3. γAb = −Abγ.4. γ2 = −I.5. Ab reverses the parity of forms.6. Abdb = −dbAb and Abd

∗b = −d∗bAb, where d

∗b = −∗db∗ is the L2–adjoint of db.

7. A2b preserves the subspace Ω

k(∂X ;E|∂X) for each k, and equals the twisted Lapla-cian on k–forms, A2

b = ∆b = dbd∗b + d∗bdb.

8. The kernel of Ab equals kerA2b = ker∆b, which is identified using the Hodge theo-

rem with the DeRham cohomology of the complex (Ωk(∂X ;E|∂X), db). The DeR-ham isomorphism identifies the DeRham cohomology with twisted cohomologyH∗(∂X ;Cn

α), where α : π1∂X → U(n) ⊂ GL(Cn) is the holonomy representationof the flat connection b.

42 PAUL KIRK AND MATTHIAS LESCH

The first 5 facts do not depend on B being a flat connection, and hold for any U(n)connection in temporal gauge near the boundary. The last three depend on b being flat.

For convenience we simplify the notation as follows. Let ΩevenX denote ⊕pΩ

2p(X ;E)and let Ω∗

∂X denote ⊕kΩk(∂X ;E|∂X). The L2 completion of Ω∗

∂X will be denoted byL2(Ω∗

∂X). We will often drop the subscripts “B” and “b” and, for example, write D forDB, A for Ab, and d for dB or db.

The self–adjoint operator A induces a spectral decomposition of L2(Ω∗∂X). We denote

the µ–eigenspace of A by Eµ. Given ν ≥ 0 we will also use the notation

F+ν = spanL2ψµ | Aψµ = µψµ, µ > ν = ⊕

µ>νEµ,

F−ν = spanL2ψµ | Aψµ = µψµ, µ < −ν = ⊕

µ<−νEµ,

E+ν = ⊕

0<µ≤νEµ,

andE−

ν = ⊕−ν≤µ<0

Eµ.

Thus E−ν is the finite–dimensional span of the eigenvectors of A with eigenvalues

µ in the range −ν ≤ µ < 0, E+ν corresponds to the range 0 < µ ≤ ν (if ν = 0,

then E±ν = 0). Similarly F−

ν is the infinite–dimensional space spanned by eigenvectorswith eigenvalues µ satisfying µ < −ν, and F+

ν corresponds to µ > ν. In particularF+0 denotes the positive eigenspan and F−

0 the negative eigenspan of A. This gives anorthogonal decomposition

L2(Ω∗∂X) = F−

ν ⊕ E−ν ⊕ kerA⊕E+

ν ⊕ F+ν . (8.2)

Another orthogonal decomposition of L2(Ω∗∂X) is the Hodge decomposition:

L2(Ω∗∂X) = im d⊕ kerA⊕ im d∗. (8.3)

We introduce a notational convention: the decomposition 8.2 is compatible with theoperators d, d∗ in the sense that we have decompositions of domains:

D(d) = (F−ν ∩ D(d))⊕E−

ν ⊕ kerA⊕ E+ν ⊕ (F+

ν ∩ D(d)),

D(d∗) = (F−ν ∩ D(d∗))⊕ E−

ν ⊕ kerA⊕E+ν ⊕ (F+

ν ∩ D(d∗).(8.4)

Note that E−ν ⊕ kerA ⊕ E+

ν consists of smooth sections hence (E−ν ⊕ kerA ⊕ E+

ν ) ∩D(d) ∩ D(d∗) = E−

ν ⊕ kerA ⊕ E+ν . By slight abuse of notation we will write in the

sequel d(∗)(F±ν ) for the image of d(∗) on F±

ν ∩ D(d(∗)).The relations between the decompositions (8.2) and (8.3) are summarized in the

following useful lemma.

Lemma 8.1.

1. d(F±ν ) ⊂ F∓

ν and d∗(F±ν ) ⊂ F∓

ν .2. F+

ν = d(F−ν )⊕ d∗(F−

ν ) = (ker d : F+ν → F−

ν )⊕ (ker d∗ : F+ν → F−

ν ).3. F−

ν = d(F+ν )⊕ d∗(F+

ν ) = (ker d : F−ν → F+

ν )⊕ (ker d∗ : F−ν → F+

ν ).4. d(E−µ) ⊂ Eµ and d∗(E−µ) ⊂ Eµ, and for µ 6= 0, Eµ = d(E−µ)⊕ d∗(E−µ).5. γ(ker d) = ker d∗ and γ(ker d∗) = ker d.

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 43

Proof. If Aβ = µβ, then Adβ = −dAβ = −µdβ, and similarly Ad∗β = −µd∗β.This proves the first assertion and the first part of the fourth assertion.

If β ∈ F+ν , then β is orthogonal to kerA, since the decomposition (8.2) is orthogonal.

Since the decomposition (8.3) is also orthogonal, β has the orthogonal decompositionβ = dτ + d∗σ. Write τ = τ− + τ+ ∈ F−

ν ⊕ F+ν , and similarly σ = σ− + σ+. Then

β = dτ− + dτ+ + d∗σ− + d∗σ+.

Since β ∈ F+ν , the first assertion implies that dτ+ = 0 = d∗σ+, so that β = dτ− + d∗σ−.

The second assertion follows from this and the consequence d(F−ν ⊕ F+

ν ) = ker d :F−ν ⊕F+

ν → F−ν ⊕F+

ν of the DeRham theorem. The third assertion is proved similarly,as is the second part of the fourth assertion.

The last assertion follows from the identity d∗ = −∗d∗ and the fact that γ equals ∗up to a non–zero constant.

Of particular concern will be the symplectic structure on kerA. The isomorphismγ preserves kerA, satisfies γ2 = −I, and acts with signature zero, since (∂X,A)bounds (X,D). Therefore kerA is a finite–dimensional Hermitian symplectic subspaceof L2(Ω∗

∂X).Notice that the restrictions of 〈 , 〉, γ, and ω to kerA induce these structures on the

cohomology H∗(∂X ;Cnα) via the Hodge and DeRham isomorphisms. The inner product

〈 , 〉 and complex structure γ depend on the choice of Riemannian metric on ∂X , butthe symplectic structure ω does not: if β1 ∈ kerA is a p-form and β2 ∈ kerA is a 2n−pform, then

ω(β1, β2) = 〈β1, γ(β2)〉 = ir∫

∂X

β1 ∧ β2 (8.5)

where the constant ir depends only on p and n (and we have suppressed the notation forthe inner product in the flat Cn bundle E|∂X). Since wedge products and cup productscorrespond via the DeRham isomorphism, ω coincides with the cup product up to apower of i, and in particular is a homotopy invariant. To put this differently, The cupproduct, together with the standard U(n)–invariant Hermitian inner product on Cn,induces a skew–hermitian form

ω : H∗(∂X ;Cn)×H∗(∂X ;Cn) → C, ω(β1, β2) = ir(β1 ∪ β2

)∩ [∂X ]

which is a homotopy invariant of the pair (∂X, α|∂X). Fixing a Riemannian metric on∂X induces a positive definite Hermitian inner product and an isomorphism γ on kerA.The Hodge and DeRham theorems define an isomorphism kerA → H∗(∂X ;Cn

α) whichtakes the form 〈x, γ(y)〉 to the form ω(x, y).

The following lemma collects some useful information about symplectic subspacesand symplectic reduction. For more details about symplectic reduction in this settingthe reader should consult Section 6.3 and [26].

Lemma 8.2.

1. Let S ⊂ L2(Ω∗∂X) be a closed subspace satisfying γ(S) ⊥ S. Then S ⊕ γ(S) is

a Hermitian symplectic subspace of L2(Ω∗∂X), and S is a Lagrangian subspace of

S ⊕ γ(S).

44 PAUL KIRK AND MATTHIAS LESCH

2. If ν ≥ 0, then F−ν ⊕ F+

ν , E−ν ⊕ E+

ν , E−ν ⊕ kerA ⊕ E+

ν , and d(E±ν ) ⊕ d∗(E∓

ν ) areHermitian symplectic subspaces of L2(Ω∗

∂X).3. Given a Lagrangian subspace L ⊂ L2(Ω∗

∂X) so that (L, F−0 ) form a Fredholm pair

of subspaces, then

Rν(L) :=L ∩ (F−

ν ⊕ E−ν ⊕ kerA⊕E+

ν )

L ∩ F−ν

⊂ E−ν ⊕ kerA⊕E+

ν (8.6)

is a Lagrangian subspace, called the symplectic reduction of L with respect toF−ν .

Proof. 1. Notice that γ preserves S ⊕ γ(S). Let K±i denote the ±i eigenspacesof γ acting on S ⊕ γ(S). It is easy to check that since γ(S) is orthogonal to S, theprojections S⊕ γ(S) → K±i restrict to isomorphisms on S. Thus the ±i eigenspaces ofγ on S⊕γ(S) have the same dimension (or are both infinite). This shows that S⊕γ(S)is a symplectic subspace of L2(Ω∗

∂X). Clearly S is a Lagrangian subspace of S ⊕ γ(S).2. For F−

ν ⊕ F+ν , take S = F−

ν and apply the first assertion. For E−ν ⊕ E+

ν ,take S = E−

ν . For d(E±ν ) ⊕ d∗(E∓

ν ), take S = d(E±ν ); then γ(S) = ∗S = ∗d(E±

ν ) =∗d(∗E∓

ν ) = d∗(E∓ν ). That kerA is symplectic was discussed above; hence the direct

sum E−ν ⊕ kerA⊕ E+

ν is symplectic.3. We apply Proposition 6.12 with U = F−

ν ⊕E−ν ⊕kerA⊕E+

ν . We have Ann(U) =F−ν and U ∩γU = E−

ν ⊕kerA⊕E+ν . Since (L, F

−0 ) form a Fredholm pair and F−

0 /F−ν is

finite–dimensional, also (L,Ann(U)) = (L, F−ν ) is Fredholm. Consequently L+Ann(U)

is closed and we reach the desired conclusion using Proposition 6.12.

In preparation for what follows we define the following enlargements of X . Givenr ≥ 0 define

Xr = ([−r, 0]× ∂X) ∪X

and

X∞ = ((−∞, 0]× ∂X) ∪X.

Thus Xr has a collar of length r attached to X and X∞ is obtained from X by attachingan infinitely long collar. Equation (8.1) can be used on the collar to define a naturalextension of D to Xr and X∞.

The key to identifying the adiabatic limit of the Calderon projector is the followingresult.

Proposition 8.3. Suppose that the boundary of X is non–empty, and suppose thatβ ∈ Ωeven

X satisfies Dβ = 0 and r(β) ∈ F−0 ⊕ kerA = spanψµ | µ ≤ 0. Then dβ = 0,

d(∗β) = 0, and d(r(β)) = 0.

Proof. Naturality of the exterior derivative implies that d(i∗(z)) = i∗(dz) for anyz ∈ Ωk

X . It suffices, therefore, to show that dβ = 0 and d(∗β) = 0, since r(β) =i∗(β) + i∗(∗β) and hence

d(r(β)) = d(i∗(β) + i∗(∗β)) = i∗(dβ + d ∗ β) = 0.

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 45

Following [1], since Dβ = 0 and r(β) ∈ F−0 ⊕ kerA, β has a Fourier expansion on

the collar [0, ε)× ∂X of the form

β|[0,ε)×∂X =∑

µ<0

cµe−xµψµ + k, (8.7)

where k ∈ kerA, x ∈ [0, ε), and ψµ ∈ Eµ. Equation 8.7 can be used to extend β to abounded form on X∞ so that the extension still satisfies Dβ = 0.

Notice that dk = 0 since k ∈ kerA and k is independent of the collar parameter.Thus dβ decays exponentially on the infinite collar (−∞, 0] × ∂X . Write β =

∑β2p.

Then

〈d ∗ β2p, ∗dβ2(p−1)〉L2(Ω∗Xr

) = ±

Xr

d ∗ β2p ∧ β2(p−1)

= ±

Xr

d(∗β2p ∧ dβ2(p−1))

= ±

∂X×−r

i∗(∗β2p) ∧ i∗(dβ2(p−1)).

The last step follows from Stokes’s theorem. As r increases to infinity, the last integralconverges to zero since ∗β2p is bounded on X∞ and dβ2(p−1) exponentially decays. It

follows that d ∗ β2p and ∗dβ2(p−1) are orthogonal in L2(Ω2(n−p−1)X∞

). Now

0 = Dβ = in+1∑

p

(−1)p(d ∗ β2p + ∗dβ2(p−1)),

with this sum expressed as a sum of homogeneous components. Thus d ∗ β2p and∗dβ2(p−1) both vanish for each p, and therefore d ∗ β and ∗dβ both vanish.

As an application of Proposition 8.3 we can identify the limiting values of extendedL2 solutions of Dβ = 0 in the sense of [1]. Recall that this is the subspace of kerAdefined by

Vα =k∣∣∣ there exists a β ∈ Ωeven

X with Dβ = 0and r(β) = f− + k ∈ F−

0 ⊕ kerA

. (8.8)

The terminology is justified by the Fourier expansion (8.7). In light of the uniquecontinuation property for D (which says that for each ℓ ∈ LX there exists a unique βwith Dβ = 0 and r(β) = ℓ), it is easy to see that Vα has the alternative description asa symplectic reduction:

Vα = R0(LX) =LX ∩ (F−

0 ⊕ kerA)

LX ∩ F−0

⊂ kerA. (8.9)

Equation (8.9) says that Vα is the symplectic reduction of the Cauchy data spaceLX with respect to subspace F−

0 . Using Lemma 8.2 it follows that Vα is a Lagrangiansubspace of kerA.

46 PAUL KIRK AND MATTHIAS LESCH

The kernel of A is identified via the Hodge and DeRham theorems with the coho-mology H∗(∂X,Cn

α). The next result identifies Vα.

Corollary 8.4. The space Vα of limiting values of extended L2 solutions toDβ = 0 on X∞ is identified via the Hodge and DeRham theorems with the image of thecohomology of X in the cohomology of ∂X (with local coefficients in the correspondingflat Cn bundle):

Vα = im i∗ : H∗(X ;Cnα) → H∗(∂X ;Cn

α).

Proof. Proposition 8.3 shows that if β ∈ ΩevenX satisfies Dβ = 0 and r(β) ∈

F−0 ⊕kerA, then β and ∗β are closed forms. Thus they represent classes in H∗(X ;Cn

α).Since r(β) = i∗(β) + i∗(∗β), it follows that r(β) is a closed form on ∂X representinga class in im i∗ : H∗(X ;Cn

α) → H∗(∂X ;Cnα). The identification of cohomology with

harmonic forms takes [r(β)] = [f− + k] to k and so

Vα ⊂ im i∗ : H∗(X ;Cnα) → H∗(∂X ;Cn

α).

The space Vα is a Lagrangian subspace, as is im i∗ : H∗(X ;Cnα) → H∗(∂X ;Cn

α) by astandard argument using Poincare duality. Since any two Lagrangian subspaces of afinite–dimensional symplectic vector space have the same dimension, Vα = im i∗.

It follows from Lemma 8.1 that E±ν = d(E∓

ν ) ⊕ d∗(E∓ν ), and so the decomposition

8.2 can be refined to

L2(Ω∗∂X) = F−

ν ⊕ d(E+ν )⊕ d∗(E+

ν )⊕ kerA⊕ d(E−ν )⊕ d∗(E−

ν )⊕ F+ν . (8.10)

The terms in this decomposition are arranged according to increasing eigenvalues. Wewill find it convenient to rewrite this in a different order, as a symplectic direct sum ofsymplectic subspaces:

L2(Ω∗∂X) = (F−

ν ⊕ F+ν )⊕ (d(E+

ν )⊕ d∗(E−ν ))⊕ (d∗(E+

ν )⊕ d(E−ν ))⊕ kerA. (8.11)

We will refer to the decomposition (8.11) frequently. Notice that F−ν ⊕F+

ν is infinite–dimensional and the other three symplectic summands in this decomposition have finitedimension.

There exists a ν ≥ 0 so that the Cauchy data space LX of D is transverse to F−ν .

This is because LX ∩ F−0 is finite–dimensional, and as ν increases, LX ∩ F−

ν decreasesto zero. Nicolaescu calls the smallest such ν the non–resonance level for D.

We can now state and prove a theorem identifying the limit of the Calderon projec-tors of D acting on Xr as r goes to infinity. Denote by Lr

X the Cauchy data space (i.e.the image of the Calderon projector) of D acting on Xr = ([−r, 0]× ∂X) ∪X .

Theorem 8.5. Let X be an odd–dimensional manifold with boundary and D theodd signature operator coupled to a flat connection B acting on X as above. Let ν ≥ 0be any number greater than or equal to the non–resonance level for D.

Then there exists a subspace

Wα ⊂ d(E+ν ) ⊂ F−

0

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 47

isomorphic to the image of

Heven(X, ∂X ;Cnα) → Heven(X ;Cn

α)

so that if W⊥α denotes the orthogonal complement of Wα in d(E+

ν ), then with respectto the decomposition (8.11) of L2(Ω∗

∂X) into symplectic subspaces, the adiabatic limit ofthe Cauchy data spaces decomposes as a direct sum of Lagrangian subspaces:

limr→∞

LrX = F+

ν ⊕ (Wα ⊕ γ(W⊥α ))⊕ d(E−

ν )⊕ Vα. (8.12)

where Vα ⊂ kerA = H∗(∂X ;Cnα) denotes the image of H∗(X ;Cn

α) → H∗(∂X ;Cnα).

Proof. Lemma 8.2 shows that the finite–dimensional vector space E−ν ⊕kerA⊕E+

ν

is a symplectic subspace of L2(Ω∗∂X).

Let Rν(LX) ⊂ E−ν ⊕ kerA⊕ E+

ν be the symplectic reduction of LX with respect tothe isotropic subspace F−

ν as in Lemma 8.2:

Rν(LX) =LX ∩ (F−

ν ⊕ E−ν ⊕ kerA⊕E+

ν )

LX ∩ F−ν

= projE−ν ⊕kerA⊕E+

ν

(LX ∩ (F−

ν ⊕ E−ν ⊕ kerA⊕E+

ν )).

Then Rν(LX) is a Lagrangian subspace of E−ν ⊕ kerA⊕E+

ν .Nicolaescu’s theorem [26, Theorem 4.9] says

limr→∞

LrX =

(limr→∞

erARν(LX))⊕ F+

ν (8.13)

(The sign in the exponent erA differs from [26] because in that paper the collar of Xr isparameterized as ∂X× [0, r].) Thus we need only to identify the limit of erARν(LX). Tohelp with the rest of the argument the reader should observe that the dynamics of erA

favor the vectors with a non–zero component in eigenspaces corresponding to positiveeigenvalues.

Let µ1 < µ2 < · · · < µq denote the complete list of eigenvalues of A in the range[−ν, ν]. Thus

E−ν ⊕ kerA⊕ E+

ν = Eµ1⊕ Eµ2

⊕ · · · ⊕ Eµq.

Given ℓ ∈ Rν(LX), we use this decomposition to write

ℓ = (ℓ1, ℓ2, · · · , ℓq).

Let µ(ℓ) denote the largest µi so that ℓi is non–zero (and hence ℓµ(ℓ)+1 = · · · = ℓq =0). Then

limr→∞

erA(

1

erµ(ℓ)ℓ

)= (0, 0, · · · , 0, ℓµ(ℓ), 0, · · · , 0).

This shows that

limr→∞

erARν(LX) = Lµ1⊕ Lµ2

⊕ · · · ⊕ Lµq⊂ Eµ1

⊕ Eµ2⊕ · · · ⊕ Eµq

,

where

Lµi= projEµi

(Rν(LX) ∩ (Eµ1⊕ · · · ⊕Eµi

))

= projEµi

(LX ∩ (F−

ν ⊕Eµ1⊕ · · · ⊕Eµi

)). (8.14)

48 PAUL KIRK AND MATTHIAS LESCH

WriteL− = ⊕

µj<0Lµj

⊂ E−ν ,

L0 = ⊕µj=0

Lµj⊂ kerA,

andL+ = ⊕

µj>0Lµj

⊂ E−ν ,

so thatlimr→∞

erARν(LX) = L− ⊕ L0 ⊕ L+ ⊂ E−ν ⊕ kerA⊕ E+

ν .

Set

Wα := L−. (8.15)

Lemma 8.6.

1. L0 = Vα.2. The spaces

(a) Wα = L−,(b) the image of Heven(X, ∂X ;Cn

α) → Heven(X ;Cnα),

(c) LX ∩ F−0 , and

(d) the L2 solutions of Dx = 0 on X∞

are all isomorphic.3. L− ⊂ d(E+

ν ).

Assuming these three facts, the rest of the proof of Theorem 8.5 is completed asfollows.

Note thatWα ⊂ d(E+ν ) ⊂ d(E+

ν )⊕d∗(E+

ν ) = E−ν . We defineW⊥

α to be the orthogonalcomplement of Wα in d(E+

ν ). Since

Wα ⊕ L+ = L− ⊕ L+ ⊂ E−ν ⊕ E+

ν =(d(E+

ν )⊕ d∗(E+ν )

)⊕

(d(E−

ν )⊕ d∗(E−ν )

)

is a Lagrangian subspace (obtained by modding out L0 and kerA), it follows fromLemma 8.1 that

L+ = d(E−ν )⊕ γ(W⊥

α ) ⊂ d(E−ν )⊕ d∗(E−

ν ),

completing the proof of Theorem 8.5.

Proof of Lemma 8.6. The first assertion follows immediately by comparing Equa-tions 8.9 and 8.14.

For the second assertion, Equation (8.14) shows that ifm ∈ L−, there exists a µi < 0and an

ℓ = (f, ℓµ1, ℓµ2

· · · , ℓµi) ∈ (F−

ν ⊕Eµ1⊕ Eµ2

⊕ · · · ⊕ Eµi) ∩ LX

with m = ℓµi. This sets up an identification of L− with F−

0 ∩LX . The unique continu-ation property identifies this (via the restriction map r) with the kernel of D with P≥0

boundary conditions, which, by Proposition 8.3 and Equation 8.7 (with k = 0), is thesame as the space of L2 harmonic forms in Ωeven

X∞. The space of L2 harmonic p-forms is

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 49

shown to be isomorphic to the image of Hp(X, ∂X ;Cnα) → Hp(X ;Cn

α) in [1, Proposition4.9].

The third assertion also follows, since if ℓ = r(β), then Proposition 8.3 says dℓ =d(r(β)) = 0. But

0 = dℓ = df + dℓµ1+ dℓµ2

+ · · ·+ dℓµi

and since d(Eµ) ⊂ E−µ,0 = dℓµi

= dm.

Hence (since µi < 0)

m ∈ ker(d : Eµi→ E−µi

) = d(E−µi) ⊂ d(E+

ν ),

completing the proof of Lemma 8.6

Remark 8.7. Notice that W⊥α denotes the orthogonal complement to Wα in the

finite–dimensional space d(E+ν ), not in L

2(Ω∗∂X).

We adopt the following notation in the rest of this section to deal with boundaryconditions. Given a manifold with boundary X , the odd signature operator D = DB

coupled to a flat connection B on X as above, and a Lagrangian subspace L ∈ LFred,then let η(D,X ;L) denote the η–invariant of the Dirac operator D with boundaryconditions given by the orthogonal projection to L. Thus,

η(D,X ;L) := η(DprojL , X)

in the previous notation. The same notation applies to the reduced η–invariant η.In a similar manner, given appropriate Lagrangian subspaces L,M,N of a Hermitian

symplectic Hilbert space we will use τµ(L,M,N) to denote the triple index of thecorresponding projections τµ(projL, projM , projN) (cf. Section 6.2).

Suppose that B and B′ are flat connections on X in temporal gauge near ∂X suchthat the holonomy representations α, α′ : π1X → U(n) of B,B′ are conjugate. Thenthere exists a gauge transformation g so that on a collar [0, ǫ) × ∂X , g = π∗(h) for agauge transformation h on ∂X satisfying B′ = g ·B. Hence the restrictions b, b′ of B,B′

to the boundary satisfy b′ = h · b. We have

DB′ = DgB = gDBg−1,

andAb′ = Ahb = hAbh

−1.

In particular, h takes the positive (resp. negative) eigenspan of Ab to the positive(resp. negative) eigenspan of Ab′ , and gives an isomorphism kerAb → kerAb′ whichcoincides via the Hodge and DeRham theorems with the isomorphism H∗(∂X ;Cn

α) →H∗(∂X ;Cn

α′) induced by conjugating the holonomies α, α′. Thus if K ⊂ kerAb is aLagrangian subspace, the λ–eigenspace of DB with F+

0 (b) ⊕K boundary conditions issent by g to the λ-eigenspace of DB′ with F+

0 (b′)⊕ h(K) boundary conditions.Since any representation α : π1X → U(n) is the holonomy representation of a flat

connection B, we conclude that given a representation α and a Lagrangian subspaceK ⊂ H∗(∂X ;Cn

α) (recall that the symplectic structure ω on H∗(∂X ;Cnα) is defined

50 PAUL KIRK AND MATTHIAS LESCH

by the cup product), the quantity η(D,X ;F+0 ⊕ K) is unambiguously defined, i.e. it

is independent of the choice of flat connection B in temporal gauge with holonomyconjugate to α, and the Lagrangian in kerAb corresponding to K via the Hodge andDeRham theorems is well defined. Of course η(D,X ;F+

0 ⊕ K) may depend on thechoice of Riemannian metric on X .

We can now turn to the splitting problem for the η–invariant of the odd signatureoperator. As in earlier sections suppose that M = M+ ∪ M− is a closed manifolddecomposed into 2 submanifolds along a separating hypersurface N . Assume that Nhas a neighborhood isometric to N × [−1, 1]. Suppose that B is a flat connection onM in temporal gauge on N × [−1, 1].

As we have seen, because the outward normal forM+ is the inward normal forM−,the operators γ and A for M− are related to those for M+ by a change in signs. Thishas the following consequences. First, whereas the conclusion of Theorem 8.5 identifiesthe limit of the Cauchy data spaces Lr

M+ of D acting on M+r , lim

r→∞LrM+ as

F+ν ⊕ (W+,α ⊕ γ(W⊥

+,α))⊕ d(E−ν )⊕ V+,α (8.16)

(in the decomposition (8.11)) for W+,α ⊂ d(E+ν ) ⊂ F−

0 a space isomorphic to the image

im(Heven(M+, ∂M+;Cn

α) → Heven(M+;Cnα))

and V+,α ⊂ kerA a space isomorphic to

im(Heven(M+;Cn

α) → Heven(N ;Cnα)).

For M− the conclusion is that limr→∞

LrM− is

F−ν ⊕ d(E+

ν )⊕ (γ(W⊥−,α)⊕W−,α)⊕ V−,α, (8.17)

where W−,α ⊂ d(E−ν ) ⊂ F+

0 is isomorphic to the image

im(Heven(M−, ∂M−;Cn

α) → Heven(M−;Cnα))

and V−,α ⊂ kerA is a space isomorphic to

im(Heven(M−;Cn

α) → Heven(N ;Cnα)).

(We assume that ν has been chosen greater than or equal to the non-resonance levelfor D acting on both M+ and M−.)

Theorem 7.7 calculates the η–invariant ofD acting onM in terms of the η–invariantsof D on M+ and M−. Take P to be the Atiyah–Patodi–Singer boundary projectionP = P+(V ) for some Lagrangian subspace V ⊂ kerA. Then Theorem 7.7 says

η(D,M) = η(D,M+;V ⊕ F+0 ) + η(D,M−;F−

0 ⊕ γ(V ))

− τµ(I − PM−, P+(V ), PM+).(8.18)

(Recall that PM± denotes the Calderon projectors onto the Cauchy data spaces LM± .)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 51

Theorem 8.8. Let D denote the odd signature operator coupled to a flat connection.For any Lagrangian subspace V ⊂ kerA,

η(D,M) = η(D,M+;V ⊕ F+0 ) + η(D,M−;F−

0 ⊕ γ(V ))− τµ(γ(V−,α), V, V+,α),

where τµ(γ(V−,α), V, V+,α) refers to the triple index in the finite–dimensional spacekerA ∼= H∗(N ;Cn

α).The triple index τµ(γ(V−,α), V, V+,α) vanishes if V = V+,α or V = γ(V−,α) and so

η(D,M) = η(D,M+;V+,α ⊕ F+0 ) + η(D,M−;F−

0 ⊕ γ(V+,α))

= η(D,M+; γ(V−,α)⊕ F+0 ) + η(D,M−;F−

0 ⊕ V−,α).

In particular, if H∗(N,Cnα) = 0, then

η(D,M) = η(D,M+;F+0 ) + η(D,M−;F−

0 ).

The main advantage that Theorem 8.8 has over Theorem 7.7 is that the Calderonprojectors have been replaced by the Atiyah–Patodi–Singer projections.

We postpone the proof of Theorem 8.8 until after two lemmas are in place. Thebasic idea is to apply Lemma 6.10 to the paths obtained by stretching the Cauchy dataspaces to their adiabatic limits.

Lemma 8.9. W+,α ⊕W−,α ⊕ (V+,α ∩ V−,α) is isomorphic to Heven(M ;Cnα).

Proof. For each integer k let (twisted coefficients in Cnα are to be understood for

all cohomology groups)

W k± = im

(Hk(M±, ∂M±) → Hk(M±)

)= ker

(i∗± : Hk(M±) → Hk(N)

)

and letV k± = im

(i∗± : Hk(M±) → Hk(N)

).

Consider the map

Ψk : Hk(M+)⊕Hk(M−) → Hk(N), Ψk(m+, m−) = i∗+(m+)− i∗−(m−)

in the Mayer–Vietoris sequence for M = M+ ∪N M−. Then there is a short exactsequence

0 → W k+ ⊕W k

− → kerΨk β−→V k

+ ∩ V k− → 0, (8.19)

where β(m+, m−) = i∗1(m+) = i∗2(m−). Moreover, the Mayer–Vietoris sequence gives ashort exact sequence

0 → coker Ψk−1 → Hk(M) → kerΨk → 0. (8.20)

Thus

dimHk(M) = dim coker Ψk−1 + dim(V k+ ∩ V k

−) + dimW k+ + dimW k

−. (8.21)

Also

dim coker Ψk−1 = dimHk−1(N)/(V k−1+ + V k−1

− )

= dimHk−1(N)− dimV k−1+ − dimV k−1

− + dim(V k−1+ ∩ V k−1

− ).(8.22)

52 PAUL KIRK AND MATTHIAS LESCH

Combining (8.21) and (8.22) and summing up over k even one obtains∑

dimH2k(M) =∑

dim(V k+ ∩ V k

−) +∑

dimW 2k+ +

∑dimW 2k

+∑

dimH2k−1(N)−∑

dimV 2k−1+ −

∑dimV 2k−1

− .(8.23)

The symplectic space H∗(N) decomposes as a symplectic sum Heven(N)⊕Hodd(N)(one way to see this is to notice that ∗ and hence γ preserves the parity of a harmonicform since N is 2n–dimensional). The Lagrangian subspace V ∗

+ =∑V k+ decomposes

accordingly into a sum of Lagrangian subspaces ⊕V 2k+ ⊕ V 2k−1

+ . Hence dim(⊕V 2k−1+ ) =

12dimHodd(N). Similarly dim(⊕V 2k−1

− ) = 12dimHodd(N). Thus the last three terms in

(8.23) cancel. Since W±,α = ⊕W 2k± and V±,α = ⊕V k

± ,

dimHeven(M) = dim(V+,α ∩ V−,α) + dimW+,α + dimW−,α,

completing the proof of Lemma 8.9.

Lemma 8.10. Let V ⊂ kerA be a Lagrangian subspace. Let LrM± denote the Cauchy

data space for D acting on M±r = ([−r, 0]×N) ∪M± when r <∞ and let L∞

M± be theadiabatic limit lim

r→∞LrM± which was identified in Theorem 8.5.

1. The dimension of the intersection LrM− ∩ Lr

M+ is independent of r ∈ [0,∞].2. The dimension of the intersection Lr

M− ∩ (F+0 ⊕ V ) is independent of r ∈ [0,∞].

3. The dimension of the intersection (F−0 ⊕γ(V ))∩Lr

M+ is independent of r ∈ [0,∞].4. τµ(γ(L

∞M−), F

+0 ⊕ V, L∞

M+) = τµ(γ(V−,α), V, V+,α).

Proof. 1. For r < ∞, the intersection LrM− ∩ Lr

M+ is isomorphic to the kernel ofDB acting on the closed manifold Mr = M+

r ∪M−r obtained by stretching the collar

of N . But this kernel is a homotopy invariant, isomorphic to Heven(M ;Cnα), and in

particular its dimension is independent of r.To compute L∞

M−∩L∞M+ , we use Theorem 8.5, or, more conveniently, its consequences

(8.16) and (8.17). These show that L∞M− ∩ L∞

M+ = W+,α ⊕W−,α ⊕ (V+,α ∩ V−,α), whichby Lemma 8.9 is also isomorphic to Heven(M ;Cn

α). Notice that the full conclusion ofTheorem 8.5 is used here.

2. and 3. are proven by the same argument. We prove 3. From the definition ofV+,α (8.9) there is an exact sequence

0 → LM+ ∩ F−0 → LM+ ∩ (F−

0 + kerA) → V+,α → 0.

It follows easily that for any subspace V ⊂ kerA there is an exact sequence

0 → LM+ ∩ F−0 → LM+ ∩ (F−

0 + γ(V )) → V+,α ∩ γ(V ) → 0. (8.24)

Lemma 8.6 identifies LM+ ∩ F−0 with the image

Heven(M+, ∂M+;Cnα) → Heven(M+;Cn

α),

and with W+,α. Thus the dimension of LM+ ∩ F−0 is independent of the length of

the collar of M+. Corollary 8.4 identifies V+,α with the image of H∗(M+;Cnα) →

H∗(∂M+;Cnα), hence its intersection with γ(V ) is independent of the length of the

collar as well. Thus the middle term in the exact sequence (8.24) is isomorphic to

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 53

W+,α ⊕ (V+,α ∩ γ(V )) and in particular its dimension is independent of the length ofthe collar; this shows that (F−

0 ⊕ γ(V )) ∩ LrM+ is independent of r for r <∞.

Now consider the case when r = ∞. In the decomposition (8.11) of L2(Ω∗N ),

F−0 ⊕ γ(V ) = F−

ν ⊕ d(E+ν )⊕ d∗(E+

ν )⊕ γ(V )

and so using (8.16) shows that (F−0 ⊕ γ(V )) ∩ L∞

M+ equals(F−ν ⊕ d(E+

ν )⊕ d∗(E+ν )⊕ γ(V )

)∩(F+ν ⊕ (W+,α ⊕ γ(W⊥

+,α))⊕ d(E−ν )⊕ V+,α

)

= W+,α ⊕ (γ(V ) ∩ V+,α).

Therefore, (F−0 ⊕ γ(V )) ∩ L∞

M+ equals to W+,α ⊕ (V+,α ∩ γ(V )).4. It follows immediately from the definition that the triple index is addi-

tive in the following sense: let H,H ′ be Hermitian symplectic Hilbert spaces andP,Q,R (resp. P ′, Q′, R′) be projections in Gr(H) (resp. Gr(H ′)) such that thetriple indices τµ(P,Q,R), τµ(P

′, Q′, R′) are well–defined. Then the triple index of(P ⊕ P ′, Q ⊕ Q′, R ⊕ R′) is well–defined in the Hermitian symplectic Hilbert spaceH ⊕H ′ and we have

τµ(P ⊕ P ′, Q⊕Q′, R⊕ R′) = τµ(P,Q,R) + τµ(P′, Q′, R′).

In the decomposition (8.11) we have

γ(L∞M−) = γ(F−

ν ⊕ d(E+ν )⊕

(γ(W⊥

−,α)⊕W−,α

)⊕ V−,α)

= F+ν ⊕ d∗(E−

ν )⊕(W⊥

−,α ⊕ γ(W−,α))⊕ γ(V−,α),

F+0 ⊕ V = F+

ν ⊕ d∗(E−ν )⊕ d(E−

ν )⊕ V, and

L∞M+ = F+

ν ⊕(W+,α ⊕ γ(W⊥

+,α))⊕ d(E−

ν )⊕ V+,α.

Using the additivity of τµ we see that τµ(γ(L∞M−), F

+0 ⊕ V, L∞

M+) equals

τµ(F+ν , F

+ν , F

+ν )F−

ν ⊕F+ν+ τµ(d

∗(E−ν ), d

∗(E−ν ),W+,α ⊕ γ(W⊥

+,α))d(E+ν )⊕d∗(E−

ν )

+ τµ(γ(W⊥−,α)⊕W−,α, d(E

−ν ), d(E

−ν ))d∗(E+

ν )⊕d(E−ν ) + τµ(γ(V−,α), V, V+,α)kerA

which equals τµ(γ(V−,α), V, V+,α)kerA by Proposition 6.11.

Proof of Theorem 8.8. Combining Lemma 8.10 and Lemma 6.10 to the pathsof Cauchy data spaces obtained by stretching the collars ofM± to their adiabatic limit,we see that (switching from projection to Lagrangian notation)

τµ(I − PM−, P+(V ), PM+) = τµ(γ(L

∞M−), F+

0 ⊕ V, L∞M+) = τµ(γ(V−,α), V, V+,α).

It then follows from Theorem 7.7 (and in particular (8.18)) that

η(D,M) = η(DB,M+;V ⊕ F+

0 ) + η(D,M−;F−0 ⊕ γ(V ))− τµ(γ(V−,α), V, V+,α).

Proposition 6.11 shows that τµ(γ(V−,α), V, V+,α) = 0 if V = γ(V−,α) or V+,α.

54 PAUL KIRK AND MATTHIAS LESCH

Remark 8.11. The Riemannian metric on the separating hypersurface N entersinto the formula of Theorem 8.8 via the map γ : kerA → kerA, since γ equalsthe Hodge–∗ operator up to a power of i. It follows that the correction termτµ(γ(V−,α), V, V+,α) is not a homotopy invariant. In fact, suppose that varying theRiemannian metric moves the space γ(V−,α) slightly (we use the Hodge and DeRhamtheorems to identify this as a subspace of the fixed space H∗(N ;Cn

α)). Then by choosingV ⊂ H∗(N ;Cn

α) carefully so that γ(V−,α) passes through V as the metric is varied onecan change τµ(γ(V−,α), V, V+,α).

To complete our analysis of the odd signature operator, we derive a formula whichcalculates η(D,M) in terms of intrinsic invariants of the pieces and a “correction term”which only depends on finite–dimensional data, namely the subspaces V±,α ⊂ kerA.

First we define the analogue of the map Φ : Gr(A) → U (Ei, E−i) of Equation (2.7)in the finite–dimensional space kerA. We use the Lagrangian notation, so that to anyLagrangian subspace K ⊂ kerA we assign the unitary map

φ(K) : (Ei ∩ kerA) → (E−i ∩ kerA) (8.25)

by the formulaK =

x+ φ(K)(x)

∣∣ x ∈ (Ei ∩ kerA).

Theorem 8.12. For the odd signature operator D coupled to a flat connection Bwith holonomy α : π1M → U(n) acting on a split manifold M =M+ ∪N M−,

η(D,M) = η(D,M+;V+,α ⊕ F+0 ) + η(D,M−;F−

0 ⊕ V−,α)

+ dim(V+,α ∩ V−,α)−1

2πitr log(−φ(V+,α)φ(V−,α)

∗).

Remark 8.13. Corollary 8.4 implies that dim(V+,α ∩ V−,α) depends only on thehomotopy type of the triple (M,M+,M−) and the representation α : π1M → U(n).

Proof. Theorem 4.4 implies that

η(D,M+;P+(γ(V−,α))) = η(D,M+;PM+)

+ 12πi

tr log(Φ(P+(γ(V−,α)))Φ(PM+)∗

) (8.26)

and that

η(D,M+;P+(V+,α)) = η(D,M+;PM+)

+ 12πi

tr log(Φ(P+(V+,α))Φ(PM+)∗

).

(8.27)

Subtracting (8.27) from (8.26) yields

η(D,M+;P+(γ(V−,α))) = η(D,M+;P+(V+,α))

+ 12πi

tr log(Φ(P+(γ(V−,α)))Φ(PM+)∗

)− 1

2πitr log

(Φ(P+(V+,α))Φ(PM+)∗

).

(8.28)

The difference1

2πitr log

(Φ(P+(γ(V−,α)))Φ(PM+)∗

)− 1

2πitr log

(Φ(P+(V+,α))Φ(PM+)∗

)

is equal to

τµ(P+(V+,α), P

+(γ(V−,α)), PM+)− 12πi

tr log(Φ(P+(V+,α))Φ(P

+(γ(V−,α)))∗)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 55

by Lemma 6.9. Moreover, it follows easily from the definitions that

tr log(Φ(P+(V+,α))Φ(P

+(γ(V−,α)))∗)= tr log

(φ(V+,α)φ(γ(V−,α))

∗),

and since φ(γ(V )) = −φ(V ), that

tr log(φ(V+,α)φ(γ(V−,α))

∗)= tr log

(− φ(V+,α)φ(V−,α)

∗).

Hence (8.28) reduces to

η(D,M+;P+(γ(V−,α))) = η(D,M+;P+(V+,α))

+ τµ(P+(V+,α), P

+(γ(V−,α)), PM+)− 12πi

tr log(− φ(V+,α)φ(V−,α)

∗).

(8.29)

We will show that

τµ(P+(V+,α), P

+(γ(V−,α)), PM+) = dim(V+,α ∩ V−,α). (8.30)

Assuming (8.30), the proof of Theorem 8.12 is completed by combining (8.29) andTheorem 8.8, taking V = γ(V−,α).

It remains, therefore, to prove (8.30). The proof is similar to the proof of Theorem8.8. Lemma 8.10 implies that as the collar ofM+ is stretched to its adiabatic limit, thedimension of the intersection of Lr

M+ with F+0 ⊕ V+,α is independent of r ∈ [0,∞], as is

the dimension of the intersection of LrM+ with F+

0 ⊕ γ(V−,α).Lemma 6.10 then implies that τµ(P

+(V+,α), P+(γ(V−,α)), PM+) is equal to

τµ(P+(V+,α), P

+(γ(V−,α)), P∞M+). Using additivity of the triple index with respect

to the decomposition (8.11), the calculation of L∞M+ (8.16), and Proposition 6.11, we

conclude that

τµ(P+(V+,α), P

+(γ(V−,α)), P∞M+) = τµ(V+,α, γ(V−,α), V+,α).

Proposition 6.11 then implies that

τµ(V+,α, γ(V−,α), V+,α) = dim(V+,α ∩ V−,α).

It is convenient to introduce the following notation.

Definition 8.14. Let (H, 〈 , 〉, γ) be a finite-dimensional Hermitian symplecticspace (cf. Def. 2.8). Define a function of pairs of Lagrangian subspaces

mH : L (H)× L (H) → R

by

mH(V,W ) = − 1πitr log(−φ(V )φ(W )∗) + dim(V ∩W )

= − 1πi

λ∈spec(−φ(V )φ(W )∗)

λ6=−1

log λ.

Here φ(V ) is the unitary map from the +i eigenspace Ei of γ to the −i eigenspace E−i

of γ whose graph is V . (If H = 0 then set mH(V,W ) = 0.) Recall that V ∩ W isisomorphic to the −1–eigenspace of φ(γV )φ(W )∗ = −φ(V )φ(W )∗ (cf. Lemma 2.6).

56 PAUL KIRK AND MATTHIAS LESCH

The function m has been investigated before, the notation is taken from [9].Given an even-dimensional Riemannian manifold (N, g) and a representation α :

π1N → U(n), define

m(V+, V−, α, g) = mH∗(N ;Cnα)(V+, V−),

where we have used the Hodge Theorem (and hence the metric g on N) to identifyH∗(N ;Cn

α) with kerA (so that γ and hence φ make sense).

Thus Theorem 8.12 says that

η(D,M) = η(D,M+;V+,α ⊕ F+0 ) + η(D,M−;F−

0 ⊕ V−,α)

+ 12dim(V+,α ∩ V−,α) +

12m(V+,α, V−,α, α, g).

(8.31)

Using η–invariants instead of η–invariants, (8.31) can be put in the more compact form

η(D,M) = η(D,M+;V+,α ⊕ F+0 ) + η(D,M+;F−

0 ⊕ V−,α) +m(V+,α, V−,α, α, g) (8.32)

using Equation (8.24) and Lemma 8.9.The function mH(V,W ) has some useful properties which we list in the following

proposition.

Proposition 8.15.

1. mH(W,V ) = −mH(V,W ).2. mH1⊕H2

(V1 ⊕ V2,W1 ⊕W2) = mH1(V1,W1) +mH2

(V2,W2).3. If ht : H → H, t ∈ [0, 1] is a continuous family of symplectic automorphisms,

then mH(ht(V ), ht(W )) is continuous in t.

Proof. The first assertion follows immediately from the definition of mH . Thesecond assertion is clear. For the third, notice that dim(ht(V )∩ ht(W )) is independentof t, and that the −1 eigenspace of −φ(ht(V ))φ(ht(W ))∗ is isomorphic to ht(V )∩ht(W ).In particular the −1 eigenspace of −φ(ht(V ))φ(ht(W ))∗ is constant dimensional, and sot 7→ log(−φ(ht(V ))φ(ht(W ))∗) is continuous. These facts imply that mH(ht(V ), ht(W ))is continuous in t.

8.2. The Atiyah–Patodi–Singer ρα–invariant for manifolds with bound-

ary. We apply the previous results to obtain information about the Atiyah–Patodi–Singer ρα–invariant [2]. Consider two flat connections: B with holonomy α, odd signa-ture operatorDB and tangential operator Ab, and the trivial connection Θ on the bundleCn × X → X with (trivial) holonomy τ , odd signature operator DΘ, and tangentialoperator Aθ.

In expressions like η(D,X ;V ⊕ F+0 ) the notation F+

0 is to be understood as thepositive eigenspan of the tangential operator A of D and V as a Lagrangian in kerA.In particular, in a “mixed” expression like η(DB, X ;Vα⊕F+

0 )− η(DΘ, X ;Vτ ⊕F+0 ) the

reader should understand that the first F+0 refers to the positive eigenspan of Ab and

the second the positive eigenspan of Aθ. These are in general unrelated since Ab actson the bundle E|∂X and Aθ acts on the trivial bundle.

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 57

Lemma 8.16. Let X be a Riemannian manifold with boundary, whose collar is iso-metric to [0, ε)× ∂X. Let B be a flat connection on a compact manifold X in temporalgauge near the boundary with holonomy α, and let Θ denote the trivial connection, withtrivial holonomy τ : π1(X) → U(n).

Then the difference

η(DB, X ;Vα ⊕ F+0 )− η(DΘ, X ;Vτ ⊕ F+

0 ) (8.33)

depends only on the diffeomorphism type of X, the conjugacy class of the holonomyrepresentation of B and the restriction of the Riemannian metric to ∂X.

Proof. We explained above why the η–invariant depends on the flat connection Bonly through the conjugacy class of its holonomy representation.

By taking the double of X we obtain a closed Riemannian manifoldM = X∪−X =M+ ∪M− over which the connections B and Θ extend flatly.

Letting DB denote the extension to M , we know from [2] that the difference

η(DB,M)− η(DΘ,M) (8.34)

is independent of the metric on M and depends only on the conjugacy class of theholonomy representation of B (see the paragraph following this proof).

Theorem 8.8 shows that

η(DB,M)− η(DΘ,M) = η(DB,M+;Vα ⊕ F+

0 ) + η(DB,M−;F−

0 ⊕ γ(Vα))

− η(DΘ,M+;Vτ ⊕ F+

0 )− η(DΘ,M−;F−

0 ⊕ γ(Vτ )).(8.35)

Notice that by Corollary 8.4 the subspaces Vα and Vτ are independent of the Rie-mannian metric onM+. Hence solving for η(DB,M

+;Vα⊕F+0 )−η(DΘ,M

+;Vτ⊕F+0 ) in

(8.35) yields an expression which is unchanged when the Riemannian metric is alteredon the interior of X =M+.

We can now extend the definition of the Atiyah–Patodi–Singer ρα–invariant to man-ifolds with boundary. Recall that the ρα–invariant is defined in [2] for a closed manifoldM and a representation α : π1(M) → U(n) by

ρ(M,α) = η(DB)− η(DΘ)

where B is a flat connection on M with holonomy α and Θ denotes a trivial U(n)connection. It is a diffeomorphism invariant of the pair (M, [α]) where [α] denotes theconjugacy class of α. In terms of reduced η-invariants ρ(M,α) can be written:

ρ(M,α) = 2(η(DB)− η(DΘ))− dimkerDB + dimkerDΘ.

Since kerDB is isomorphic to Heven(M ;Cnα) this is the same as

ρ(M,α) = 2(η(DB,M)− η(DΘ,M))− dimHeven(M ;Cnα) + dimHeven(M ;Cn). (8.36)

Definition 8.17. Given a triple (X,α, g), where

1. X is a compact odd–dimensional manifold with boundary,2. α : π1(X) → U(n) is a representation, and3. g is a Riemannian metric on ∂X ,

58 PAUL KIRK AND MATTHIAS LESCH

choose a Riemannian metric on X isometric to [0, ǫ)× ∂X on a collar of ∂X and a flatconnection B with holonomy α in temporal gauge near the boundary. Then define

ρ(X,α, g) := η(DB, X, F+0 ⊕ Vα)− η(DΘ, X, F

+0 ⊕ Vτ ).

Reversing the orientation of X changes the sign of ρ(X,α, g), since the η–invariantchanges sign when the orientation is reversed.

In terms of reduced η invariants ρ(X,α, g) can be expressed as:

ρ(X,α, g) = 2(η(DB, X, F

+0 ⊕ Vα)− η(DΘ, X, F

+0 ⊕ Vτ )

)− dimWα + dimWτ ,

whereWα

∼= im(Heven(X, ∂X ;Cn

α) → Heven(X ;Cnα))

andWτ

∼= im(Heven(X, ∂X ;Cn) → Heven(X ;Cn)

).

This is because the kernel of DB acting on X with boundary conditions given bythe Atiyah–Patodi–Singer projection P+(Vα) is isomorphic to Wα by Lemma 8.6 andEquation (8.24) (with V = V+,α), and similarly for DΘ.

Lemma 8.16 shows that ρ(X,α, g) is independent of the choice of Riemannian metricon the interior of X (as long as the metric is a product in some collar of the boundary)and the choice of flat connection B with holonomy α.

We now can state the main result of this section.

Theorem 8.18. Suppose the closed manifold M contains a hypersurface N sep-arating M into M+ and M−. Fix a Riemannian metric g on N . Suppose thatα : π1(M) → U(n) is a representation, and let τ : π1(M) → U(n) denote the triv-ial representation.

Then

ρ(M,α) = ρ(M+, α, g) + ρ(M−, α, g) +m(V+,α, V−,α, α, g)−m(V+,τ , V−,τ , τ, g).

Proof. This follows by applying Equation (8.32) to B and Θ and subtracting.

It can be shown that the invariants ρ(M±, α, g) and m(V+, V−, α, g) depend in gen-eral on the choice of Riemannian metric g on the hypersurface N . We leave an anintriguing open problem to determine exactly how they depend on the metric g, and inparticular, how these invariants change if g is replaced by the pulled–back metric f ∗(g)for a diffeomorphism f : N → N .

8.3. Relationship to Wall’s non-additivity theorem. Theorems 8.18 and 8.12should be viewed as odd–dimensional counterparts of Wall’s non–additivity theoremfor the signature [32]. Indeed these theorems give formulas which express the “non-additivity of the signature defect”. The relationship between splitting theorems for theη–invariant and Wall non–additivity is explored in Bunke’s article [9] and also in [20].

To clarify the relationship between Wall’s theorem and Theorem 8.18, consider thefollowing situation. Suppose we are given two 4k–dimensional manifolds Z+ and Z−

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 59

with ∂Z± = M± ∪N M0. Suppose that ∂M0 = N and that Z = Z+ ∪M0 Z−. Finallysuppose that α : π1Z → U(n) is a representation and let τ : π1Z → U(n) denote thetrivial representation. The Atiyah–Patodi–Singer signature theorem [2, Theorem 2.4]says that

Signτ (Z)− Signα(Z) = ρ(M,α).

Similarly Signτ (Z+) − Signα(Z

+) = ρ(M+ ∪ M0, α) and Signτ (Z−) − Signα(Z

−) =ρ(−M0 ∪M−, α). On the other hand Wall’s theorem says that

Signα(Z) = Signα(Z+) + Signα(Z

−)− σ(V+,α, V−,α, V0,α, α),

where σ is a correction term which depends on the relative positions of the subspacesV+,α, V−,α and V0,α in H∗(N ;Cn

α). Similarly Signτ (Z) = Signτ (Z+) + Signτ (Z

−) −σ(V+,τ , V−,τ , V0,τ , τ).

Hence

σ(V+,α, V−,α, V0,α, α)− σ(V+,τ , V−,τ , V0,τ , τ)

= Signτ (Z)− Signτ (Z+)− Signτ (Z

−)

−(Signα(Z)− Signα(Z

+)− Signα(Z−))

= ρ(M+ ∪M−, α)− ρ(M+ ∪M0, α)− ρ(−M0 ∪M−, α)

(8.37)

Applying Theorem 8.18 we see that (8.37) is equal to σα − στ , where

σα := m(V+,α, V−,α, α, g)−m(V+,α, V0,α, α, g)−m(V0,α, V−,α, α, g)

and

στ := m(V+,τ , V−,τ , τ, g)−m(V+,τ , V0,τ , τ, g)−m(V0,τ , V−,τ , τ, g).

This motivates introducing the following notation. Given a Hermitian symplecticspace H , define the function of triples of Lagrangian subspaces

σH : L (H)× L (H)× L (H) → Z

by

σH(V,W, U) := mH(V,W ) +mH(W,U) +mH(U, V ). (8.38)

By definition σα = σH∗(N ;Cnα)(V+,α, V−,α, V0,α) and similarly for στ . That σH is an integer

can be seen by exponentiating and using the multiplicativity of the determinant:

exp(2πi σH(V,W, U))

=(exp(tr log(−φ(V )φ(W )∗) + tr log(−φ(W )φ(U)∗) + tr log(−φ(U)φ(V )∗)

)2

= det((−1)3φ(V )φ(W )∗φ(W )φ(U)∗φ(U)φ(V )∗)2

)

= 1.

Proposition 8.19. The function σH satisfies the following properties.

1. Given a permutation β, σH(Vβ(1), Vβ(2), Vβ(3)) = sign(β)σH(V1, V2, V3).2. σH1⊕H2

(V1 ⊕ V2,W1 ⊕W2, U1 ⊕ U2) = σH1(V1,W1, U1) + σH2

(V2,W2, U2).

60 PAUL KIRK AND MATTHIAS LESCH

3. If h : H → H is a symplectic automorphism, then σH(h(V ), h(W ), h(U)) =σH(V,W, U).

4. Take H = C2 with γ =

(0 −11 0

). Then σH(C(1, 0),C(1, 1),C(0, 1)) = 1.

Proof. The first and second assertions follow from the first and second assertionsof Proposition 8.15.

For the third assertion, we first claim that the group Sp(H) of symplectic automor-phisms of H is path connected. To see this, fix a Lagrangian subspace V of H . Themap Sp(H) → L (H) taking g to g(V ) is a fibration with fiber the subgroup SpV (H)consisting of those symplectic automorphisms which leave V invariant. Next, SpV (H)fibers over GL(V ) by mapping g ∈ SpV (H) to the restriction g|V . The fiber F of thismap consists of those symplectic transformations g so that g restricts to the identity onV . Writing H = V ⊕ γ(V ) it is easy to see that F consists of all matrices of the form

(I A0 I

)

with A an arbitrary real matrix. Thus F is contractible, and since GL(V ) is pathconnected, SpV (H) is also path connected. Finally, since Sp(H) fibers over the pathconnected space L (H) ∼= U(n) with path connected fiber SpV (H), it is itself pathconnected.

Choose a path ht from the identity matrix to h. The third assertion of Proposition8.15 shows that mH(ht(V ), ht(W )) varies continuously in t. Thus the same is true ofthe integer-valued function t 7→ σH(ht(V ), ht(W ), ht(U)). Therefore this function isconstant, completing the proof of the third assertion.

To prove the last fact, Notice that C2 = Ei ⊕ E−i, where Ei is the span of (1,−i)and E−i is the span of (1, i). It is easy to calculate that with respect to these bases,

φ(C(1, 0)) = 1, φ(C(1, 1)) = −i, and φ(C(0, 1)) = −1.

Thus

σH(C(1, 0),C(1, 1),C(0, 1)) = − 1πi(log(−1 · i) + log(−(−i) · (−1)) + log(−(−1)(1)))

= 1

It follows from Proposition 8.19 and [10, Theorem 8.1] (suitably generalized to thecomplex Hermitian case) that σH is equal to the Maslov triple index τH defined in loc.cit. Therefore, σH depends only on the underlying symplectic form ω(x, y) = 〈x, γy〉,and not on the Hermitian metric. In particular σα and στ are independent of theRiemannian metric on N .

σH and the Maslov triple index τµ defined in Section 6.2 are (of course) intimatelyrelated. τµ is, up to normalization, what Bunke [9] called the twisted Maslov triple

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 61

index. A direct calculation shows the following:

σH(V,W, U) = −τµ(V,W, U)− τµ(γV,W, U)− τµ(V, γW,U) (8.39)

− τµ(V,W, γU) + dim(V ∩W ) + dim(W ∩ U) + dim(V ∩ U),

τµ(V,W, U) =1

4

(σH(V,W, U)− σH(γV,W, U)− σH(V, γW,U) (8.40)

− σH(V,W, γU) + 2 dim(γV ∩W ) + 2 dim(W ∩ γU) + 2 dim(V ∩ γU)).

Using Proposition 8.2 of loc. cit. we conclude that σα equals Wall’s correction termσ(V+,α, V−,α, V0,α) and similarly στ equals σ(V+,τ , V−,τ , V0,τ).

Using the argument outlined in [20] one can analyze Wall’s theorem using Theorem8.12 as follows.

In the context described above, the Atiyah–Patodi–Singer signature theorem statesthat

Signα(Z) = n

Z

L− η(DB,M+ ∪M−),

where L denotes the L–polynomial of the Riemannian curvature tensor on Z. Similarlyone obtains formulas for Signα(Z

±)

Signα(Z+) = n

Z+

L− η(DB,M+ ∪M0)

and

Signα(Z−) = n

Z+

L− η(DB,−M0 ∪M−).

Applying Theorem 8.12 and using Lemma 6.9 as before, one obtains

Signα(Z)− Signα(Z+)− Signα(Z

−) = σα + n( ∫

Z

L−

Z+

L−

Z−

L).

At this point one can invoke Wall’s theorem and the identification of σα with Wall’scorrection term given above to conclude that the integrals cancel. (This is not immedi-ate since the Riemannian metrics on Z+ and Z− need to be in cylindrical form near theboundary to apply the Atiyah–Patodi–Singer theorem, but these do not glue to give asmooth metric on Z in cylindrical form near the boundary.)

On the other hand, Wall’s theorem can be proven by showing that the integralscancel. This is discussed in [20] and so one obtains an analytic proof of Wall’s theorem.

More importantly, Equation (8.38) establishes a direct relationship between thecorrection terms σα and στ for the non-additivity of the signature to the correctionterm m(V+,α, V−,α, α, g)−m(V+,τ , V−,τ , τ, g) for the non-additivity of the ρ invariant.

8.4. Adiabatic stretching and general Dirac operators. Some of the pre-ceding exposition for the odd signature operator extends to the more general contextof arbitrary Dirac operators, and we discuss aspects of this now. The new feature ofthis approach is that the role of adiabatic stretching in the splitting formula for theη–invariant is clarified.

Suppose we are given an arbitrary Dirac operator D on a split manifold M =M+ ∪N M−. Assume as usual that D = γ( d

dx+ A) on a collar of N . Let Mr denote

62 PAUL KIRK AND MATTHIAS LESCH

the manifold obtained by replacing the collar [−1, 1]× N of N by the stretched collar[−r, r] × N . Thus M0 = M . Given 0 ≤ r < ∞, let Lr

M± denote the Cauchy dataspace of the operator D acting on M±

r = M± ∪ ([−r, 0]× N), and let L∞M± denote the

adiabatic limit limr→∞

LrM± . Lemma 3.2 of [14] states that the path [0,∞] → Gr(A) given

by r 7→ LrM± is continuous.

We let F±ν denote the span of λ–eigenvectors of A for ±λ > ν, and E±

ν the span ofλ–eigenvectors of A for 0 < ±λ ≤ ν, so that the L2–sections over N decompose as

F−ν ⊕ E−

ν ⊕ kerA⊕ E+ν ⊕ F+

ν

or, as a symplectic direct sum

(F−ν ⊕ F+

ν )⊕ (E−ν ⊕E+

ν )⊕ kerA. (8.41)

Theorem 8.5 has a counterpart for general Dirac operators, but the conclusion isslightly weaker. The following theorem has a similar but simpler proof than Theorem8.5. It is implied by Theorem 6.5 of [14].

Theorem 8.20. Let V+ ⊂ kerA denote the limiting values of extended L2 solutionson M+, so V+ = projkerA(LM+ ∩ (F− ⊕ kerA)). Let ν ≥ 0 be a number in the non-resonance range of D, i.e. LM+ ∩ F−

ν = 0. Then there exists a subspace W+ ⊂ E−ν

isomorphic to the space of L2 solutions to Dβ = 0 on M+∞ so that letting W⊥

+ ⊂ E−ν

denote the orthogonal complement of W+ in E−ν ,

L∞M+ = F+

ν ⊕ (W+ ⊕ γ(W⊥+ ))⊕ V+

in the decomposition (8.41). Moreover, γ(L∞M+) ∩ LM+ = 0.

Then we have the following theorem.

Theorem 8.21. With notation as above, for any r0 ≥ 0,

η(D,Mr0)− η(D,M+;L∞M+)− η(D,M−; γ(L∞

M+))= Mas(Lr

M−, L∞M+)r∈[r0,∞] −Mas(Lr

M− , LrM+)r∈[r0,∞]

Remark 8.22. In light of Theorems 7.5 and 7.6 the term Mas(LrM−, Lr

M+)r∈[r0,∞]

in Theorem 8.21 can be thought of as the spectral flow of the family of opera-tors on M obtained by stretching the collar from r0 to infinity. Similarly the termMas(Lr

M− , L∞M+)r∈[r0,∞] can be thought of as the spectral flow of the family on M− ob-

tained by using the projection to L∞M+ as boundary conditions and stretching the collar

of M− from r0 to infinity.

Proof. We prove this for r0 = 0, the general case is obtained by reparameterizing.Let Pt denote the path of projections to the Cauchy data space Lr

M+ , where t = 1/(r+1).Applying Theorem 5.9 and Proposition 5.1 to the path Pt we see that η(D,Mr0) −η(D,M+;L∞

M+) − η(D,M−; γ(L∞M+)) equals SF(DPt

,M+) + SF(DI−Pt,M−)[0,1], which

by Theorem 7.5 equals

Mas(Pt, PM+) + Mas(PM−, I − Pt). (8.42)

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 63

Switching to Lagrangian notation and parameterizing by r instead of t we can rewrite(8.42) as

−Mas(γ(LrM+), LM+)r∈[0,∞] −Mas(LM−, Lr

M+)r∈[0,∞]. (8.43)

We use the homotopy invariance of the Maslov index to simplify these terms. Con-sider first Mas(γ(Lr

M+), LM+)r∈[0,∞]. We will show this term vanishes.The path r 7→ γ(Lr

M+) is homotopic to the composite of r 7→ γ(LrM+) and the

constant map at γ(L∞M+), and the constant path at LM+ is homotopic to the composite of

r 7→ LrM+ and its inverse. Since γ(Lr

M+)∩LrM+ = 0 for all r, Mas(γ(Lr

M+), LrM+) = 0 and

so by path additivity of the Maslov index, Mas(γ(LrM+), LM+) = −Mas(γ(L∞

M+), LrM+).

Theorem 8.20 says that γ(L∞M+) ∩ Lr

M+ = 0 for r = 0, but by reparameterizing wesee that the intersection is zero for all r < ∞; obviously γ(L∞

M+) ∩ L∞M+ = 0. Hence

Mas(γ(L∞M+), Lr

M+) = 0 and so

Mas(γ(LrM+), LM+)r∈[0,∞] = 0. (8.44)

Consider now the term Mas(LM−, LrM+)r∈[0,∞]. The constant path at LM− is homo-

topic to the composite of r 7→ LrM− and its inverse, and the path r 7→ Lr

M+ is homotopicto its composite with the constant path at L∞

M+ . Therefore,

Mas(LM− , LrM+) = Mas(Lr

M− , LrM+)−Mas(Lr

M−, L∞M+). (8.45)

Substituting (8.44) and (8.45) into (8.43) finishes the proof.

We finish this article by outlining a few ways to use Theorem 8.21 to obtain otheruseful splitting formulas for the η–invariant. We will not give an exhaustive list, butwe note that many other useful formulas can be derived from these using the results ofSections 5, 6, and 7. One can, of course, obtain other formulas by reversing the rolesof M+ and M− in Theorem 8.21 and in these examples.

Example 8.23. Suppose that L∞M− ∩L∞

M+ = 0. Then there exists an r0 ≥ 0 so thatLrM− ∩ Lr

M+ = 0 and LrM− ∩ L∞

M+ = 0 for all r ≥ r0. Applying Theorem 8.21 we seethat if r ≥ r0 then

η(D,Mr) = η(D,M+;L∞M+) + η(D,M−; γ(L∞

M+))

= η(D,M+;F+ν ⊕W+ ⊕ γ(W⊥

+ )⊕ V+) + η(D,M−;F−ν ⊕W⊥

+ ⊕ γ(W+)⊕ γ(V+)).

The hypothesis L∞M− ∩ L∞

M+ = 0 is a technically simpler replacement of the hypothesis“no exponentially small eigenvalues” which appears in related results in the literature.

Example 8.24. Suppose thatDβ = 0 has no L2 solutions onM+∞; i.e. thatW+ = 0

in Theorem 8.20. Then L∞M+ = F+ ⊕ V+ and so

η(D,M)− η(D,M+;F+ ⊕ V+)− η(D,M−;F− ⊕ γ(V+))

= Mas(LrM−, F+ ⊕ V+)r∈[0,∞] −Mas(Lr

M− , LrM+)r∈[0,∞].

In other words, with respect to the Atiyah–Patodi–Singer boundary conditions givenby the projection to F+ ⊕ V+ on M+ and the projection to F− ⊕ γ(V+) on M−, thefailure of the additivity of the η–invariants is measured by Mas(Lr

M− , F+⊕V+)r∈[0,∞]−Mas(Lr

M− , LrM+)r∈[0,∞]. As remarked above this is the difference of the spectral flow of

64 PAUL KIRK AND MATTHIAS LESCH

D on M− with F+ ⊕ V+ conditions as the collar of M− is stretched to infinity, and thespectral flow of D on M as the collar is stretched to infinity.

Example 8.25. We can combine the previous two examples as follows. Supposethat there are no L2 solutions on M+

∞ and M−∞ (i.e. W+ = 0 = W−) and that the

limiting values of extended L2 solutions are transverse (i.e. V+ ∩ V− = 0 in kerA; thishappens for example if kerA = 0). Then L∞

M± = F± ⊕ V± and so both of the previousexamples apply.

Hence there exists an r0 ≥ 0 so that LrM− ∩ Lr

M+ = 0 and LrM− ∩ L∞

M+ = 0 for allr ≥ r0. Therefore,

η(D,M)− η(D,M+;F+ ⊕ V+)− η(D,M−;F− ⊕ γ(V+))

= SF(D,M−r ;F

+ ⊕ V+)r∈[0,r0] − SF(D,Mr)r∈[0,r0].

This says that the failure of additivity of the η–invariants with Atiyah–Patodi–Singerboundary conditions is measured by the difference of the spectral flow of D on M−

with F+ ⊕ V+ conditions as the length of the collar of M− is stretched to r0, and thespectral flow of D on M as the collar is stretched to r0.

In particular, if r ≥ r0,

η(D,Mr) = η(D,M+;F+ ⊕ V+) + η(D,M−;F− ⊕ γ(V+)).

This last formula appears in Bunke’s article [9]. The reader should compare this formulawith the formula of Theorem 8.8 (with V = V+,α) which, by contrast, holds in completegenerality for the odd signature operator.

These examples, together with Theorem 8.20, underscore the point that difficultiesin establishing simple splitting formulas for the η–invariant using Atiyah–Patodi–Singerboundary conditions arise from the existence of L2 solutions on the two parts of thedecomposition of M . To put this in a positive perspective, the failure of the additiv-ity of the η–invariant with Atiyah–Patodi–Singer boundary conditions is measured bythe spectral flow terms discussed in Remark 8.22 and symplectic invariants of the La-grangian subspaces W± ⊕ γ(W⊥

± ) in the finite–dimensional symplectic space E−ν ⊕ E+

ν

consisting of the span of the µ–eigenvectors of A with −ν ≤ µ ≤ ν, µ 6= 0. In ouranalysis of the odd signature operator the formulas simplify because we can controlthese terms; the spectral flow terms vanish for topological reasons and the symplecticinvariants of the Lagrangian subspaces W± ⊕ γ(W⊥

± ) vanish because of the additionalcontrol on W± that Theorem 8.5 provides over Theorem 8.20.

References

[1] M. F. Atiyah, V. K. Patodi, and I. M. Singer: Spectral asymmetry and Riemannian geometryI. Math. Proc. Camb. Phil. Soc. 77 (1975), 43–69

[2] M. F. Atiyah, V. K. Patodi, and I. M. Singer: Spectral asymmetry and Riemannian geometryII. Math. Proc. Camb. Phil. Soc. 78 (1975), 405–432

[3] J. Avron, R. Seiler, and B. Simon: The index of a pair of projections. J. Funct. Anal. 120(1994), 220–237

[4] B. Booss-Bavnbek and K. Furutani: The Maslov index: a functional analytical definition andthe spectral flow formula. Tokyo J. Math. 21 (1998), 1–34

η–INVARIANT, MASLOV INDEX, AND SPECTRAL FLOW 65

[5] B. Booß-Bavnbek and K. P. Wojciechowski: Elliptic Boundary Problems for Dirac Opera-tors. Birkhauser, Basel (1993)

[6] J. Bruning and M. Lesch: On boundary value problems for Dirac type operators: I. Regularityand self–adjointness. Preprint, 55p. (1999). Math.FA/9905181

[7] J. Bruning and M. Lesch: On the eta–invariant of certain non–local boundary value problems.Duke Math. J. 96 (1999), 425–468. Dg-ga/9609001

[8] J. Bruning and M. Lesch: Spectral theory of boundary value problems for Dirac type operators.In: B. Booß-Bavnbek and K. P. Wojciechowski (eds.), Geometric Aspects of PDE - SpectralInvariants (Roskilde Conf. Sept 1998), vol. 242 of Contemp. Math. AMS, Providence, RI (1999),pp. 203–215. Math.DG/9902100

[9] U. Bunke: On the gluing problem for the η–invariant. J. Diff. Geom. 41 (1995), 397–448[10] S. Cappell, R. Lee, and E. Miller: On the Maslov index. Comm. Pure Appl. Math. 47 (1994),

121–186[11] S. Cappell, R. Lee, and E. Miller: Self-adjoint elliptic operators and manifold decompositions

II: Spectral flow and Maslov index. Comm. Pure Appl. Math. 49 (1996), 869–909[12] X. Dai and D. S. Freed: η–invariants and determinant lines. J. Math. Phys. 35 (1994), 5155–

5194[13] M. Daniel: An extension of a theorem of Nicolaescu on spectral flow and the Maslov index. Proc.

Amer. Math. Soc. 128 (2000), no. 2, 611–619[14] M. Daniel and P. Kirk: A general splitting formula for the spectral flow (with an appendix by

K. P. Wojciechowski). Mich. Math. J. 46 (1999), 589–617[15] R. G. Douglas and K. P. Wojciechowski: Adiabatic limits of the η-invariants the odd–

dimensional Atiyah–Patodi–Singer problem. Commun. Math. Phys. 142 (1991), 139–168[16] G. Grubb: Poles of zeta and eta functions for perturbations of the Atiyah-Patodi-Singer problem.

University of Copenhagen, Preprint No. 14 (1999)[17] G. Grubb: Trace expansions for pseudodifferential boundary problems for Dirac–type operators

and more general systems. Arkiv f. Matematik 37 (1999), 45–86[18] G. Grubb and R. Seeley: Weakly parametric pseudodifferential operators and Atiyah–Patodi–

Singer boundary problems. Invent. Math. 121 (1995), 481–529[19] G. Grubb and R. T. Seeley: Zeta and eta functions for Atiyah–Patodi–Singer operators. J.

Geom. Anal. 6 (1996), 31–77[20] A. Hassell, R. Mazzeo, and R. Melrose: A signature formula for manifolds with corners of

codimension two. Topology 36 (1997), no. 5, 1055–1075.[21] T. Kato: Perturbation theory for linear operators, vol. 132 of Grundlehren der Mathematischen

Wissenschaften. 2nd edn. Springer–Verlag, Berlin–Heidelberg–New York (1976)[22] N. H. Kuiper: The homotopy type of the unitary group of Hilbert space. Topology 3 (1965),

19–30[23] M. Lesch: Determinants of Dirac type operators on manifolds with boundary. In preparation[24] M. Lesch and K. P. Wojciechowski: On the η–invariant of generalized Atiyah–Patodi–Singer

boundary value problems. Ill. J. Math. 40 (1996), 30–46[25] W. Muller: On the L2–index of Dirac operators on manifolds with corners of codimension two.

I. J. Diff. Geom. 44 (1996), 97–177[26] L. Nicolaescu: The Maslov index, the spectral flow, and decompositions of manifolds. Duke

Math. J. 80 (1995), 485–533[27] R. S. Palais: Seminar on the Atiyah-Singer Index Theorem. Annals of Mathematics Studies,

No. 57 Princeton University Press, Princeton, N.J. 1965 x+366 pp.[28] J. Phillips: Self-adjoint Fredholm operators and spectral flow. Canad. Math. Bull. 39 (1996),

460–467[29] S. G. Scott: Determinants of Dirac boundary value problems over odd-dimensional manifolds.

Commun. Math. Phys. 173 (1995), 43–76

66 PAUL KIRK AND MATTHIAS LESCH

[30] S. G. Scott and K. P. Wojciechowski: The ζ–determinant and Quillen determinant for aDirac operator on a manifold with boundary. Preprint, IUPUI (1999)

[31] R. T. Seeley: Topics in pseudo–differential operators. C.I.M.E., Conference on pseudo–differential operators 1968, Edizioni Cremonese, Roma, 1969, pp. 169–305

[32] C. T. C. Wall: Non-additivity of the signature. Invent. Math. 7 (1969), 269–274[33] K. P. Wojciechowski: The additivity of the η–invariant. The case of an invertible tangential

operator. Houston J. Math. 20 (1994), 603–621[34] K. P. Wojciechowski: The additivity of the η–invariant. The case of a singular tangential

operator. Commun. Math. Phys. 109 (1995), 315–327[35] K. P. Wojciechowski: The ζ-determinant and the additivity of the η-invariant on the smooth,

self-adjoint Grassmannian. Commun. Math. Phys. 201 (1999), 423–444

Department of Mathematics, Indiana University, Bloomington, IN, 47405, USA

E-mail address : [email protected]: http://php.indiana.edu/∼pkirk

The University of Arizona, Department of Mathematics, 617 N. Santa Rita, Tucson,

AZ, 85721–0089, USA

E-mail address : [email protected]: http://www.math.arizona.edu/∼lesch